Download The operator hierarchy, a chain of closures linking matter, life

Document related concepts

Systems theory wikipedia , lookup

Reductionism wikipedia , lookup

Introduction to evolution wikipedia , lookup

Evolving digital ecological networks wikipedia , lookup

Fundamental interaction wikipedia , lookup

Hologenome theory of evolution wikipedia , lookup

Evolutionary history of life wikipedia , lookup

Precambrian body plans wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Incomplete Nature wikipedia , lookup

Transcript
$/7(55$6&,(17,),&&2175,%87,216
Gerard Jagers op Akkerhuis
Gerard Jagers op Akkerhuis
2PVODJ'LVVHUWDWLHGHILQLWLHI*HUDUG-DJHUVLQGG
A chain of closures linking matter,
life and artificial intelligence
$/7(55$6&,(17,),&&2175,%87,216
The Operator Hierarchy
This thesis presents a new theory: the Operator
Hierarchy. The theory offers an original approach
to understanding evolution; it uses cyclical selforganization patterns to characterize all ‘particles’ in nature from hadrons and atoms to living
organisms. In practice, the Operator Hierarchy
allows novel solutions to fundamental questions,
such as the definition of life and the next step in
evolution.
The Operator Hierarchy
The operator hierarchy
A chain of closures linking matter, life and artificial intelligence
Gerard Jagers op Akkerhuis
2
Abstract
The operator hierarchy
A chain of closures linking matter, life and artificial intelligence
Een wetenschappelijke proeve op het gebied van de Natuurwetenschappen,
Wiskunde en Informatica
PROEFSCHRIFT
Ter verkrijging van de graad van doctor
aan de Radboud Universiteit Nijmegen
op gezag van de rector magnificus prof. mr. S.C.J.J. Kortmann,
volgens besluit van het college van decanen
in het openbaar te verdedigen op
maandag 6 september 2010
om 15.30 uur precies
door
Gerhardus Adrianus Johannes Maria Jagers op Akkerhuis
Geboren op 21 maart 1960
Te Hengelo (Overijssel)
4
Abstract
Promotor:
Prof. dr. H. Zwart
Manuscriptcommissie:
Prof. dr. H. Barendregt
Prof. dr. F.P. Heylighen (VU België)
Prof. dr. N.M. van Straalen (VU)
Paranimfen:
Hans-Peter Koelewijn
Leen G. Moraal
ISBN:
Lay-out:
978-90-327-0389-9
Mitekst Service v.o.f.
© 2010 G.A.J.M. Jagers op Akkerhuis; all rights reserved.
Jagers op Akkerhuis, G.A.J.M. (2010) The operator hierarchy: a chain of
closures linking matter, life and artificial intelligence. PhD thesis, Radboud
University Nijmegen.
The research presented in this PhD thesis was financially supported by the
Netherlands Integrated Soil Research Programme, the National Environmental
Research Institute (Silkeborg, Denmark), Alterra Wageningen UR and the
Radboud University Nijmegen.
Contents
Abstract
7
1.
General Introduction
9
2.
The birth of an idea
63
3.
The future of evolution
77
4.
A backbone for system science
99
5.
Bringing the definition of life to closure
121
6.
Getting the biggest part of the pie
149
7.
Definitions behind the operator theory
161
8.
Centripetal science
173
9.
General discussion and conclusions
199
References
215
Summary
231
Samenvatting
243
History and acknowledgements
255
Curriculum vitae
263
Glossary
271
Index
283
Abstract
7
Abstract
Jagers op Akkerhuis, G.A.J.M., (2010). The operator hierarchy: a chain of
closures linking matter, life and artificial intelligence
PhD thesis, Radboud University Nijmegen.
Science suffers from a knowledge explosion that is associated with
overspecialisation and compartmentalisation. To overcome this crisis, scientists
face the challenge of constructing comprehensive and integrating theory. In
relation to this challenge, this thesis presents a general theory for recognising
hierarchy in the organisation of nature. Novel about this theory is that it unifies in
a single, strict ranking the two system types that traditionally are regarded as
‘physical particles’ and as ‘organisms’. This ranking is made possible by
focusing on situations in which systems that show cyclical self-organisation
create a next system type with this property as the first possibility nature allows.
Due to this focus the result is not just a classification but a chain of systems
types resulting from and linked by process steps. This chain includes
fundamental particles, hadrons, atoms, molecules, prokaryote cells, eukaryote
cells, prokaryote and eukaryote multicellulars, and animals. The proposed theory
refers, in a general way, to the latter systems as ‘operators’ and to their
hierarchical ranking as the ‘operator hierarchy’.
After discussing in detail the construction of the operator hierarchy, it is shown
how the operator theory contributes to fundamental unification topics in science.
Applications in this field include the definitions of life, the organism and death,
the prediction of future operators, a periodic table for periodic tables, the
supplementation of evolutionary theory with the opportunity to focus on structural
development, and a discussion of the need to redefine the meme concept. In the
discussion section the theory’s strengths and weaknesses are discussed. It is
concluded that the operator theory represents a fundamental contribution to
science that offers new roads towards solving integration problems.
Keywords: evolution, operator hierarchy, natural philosophy, major evolutionary
transitions, particle hierarchy, closure, memes, arrow of complexity, hierarchy
theory, system theory, exobiology, artificial intelligence, periodic table, definition
of life, Big History.
Correspondence: Gerard Jagers op Akkerhuis,
Alterra, Wageningen UR.
P.O. Box 47, NL-6700 AA Wageningen, The Netherlands
Email: [email protected]
8
Abstract
General introduction
Chapter
9
1
General Introduction
Now, in order to work hard on something, you have to get yourself believing that
the answer’s over there, so you’ll dig hard there, right? So you temporarily
prejudice or predispose yourself – but all the time, in the back of your mind,
you’re laughing.
(Richard Feynman)
Read, every day, something no one else is reading.
Think, every day, something no one else is thinking.
Do, every day, something no one else would be silly enough to do.
It is bad for the mind to continually be part of unanimity.
(Christopher Morley)
To all those who have inspired me
Omnia lusus naturae
10
1
Chapter 1
General introduction
Science suffers from a knowledge explosion (e.g., Sagasti, 1999). In the last 30
years more discoveries have been made than ever before (Linowes, 1990). In
what could be regarded as a centrifugal vortex of scientific discovery, we have
created tools that have helped us to investigate the world we live in, which has
increased our understanding of the world, which has been used to create better
tools that have helped us to expand our understanding further still. These
developments have given us insight into the miniscule and the enormous, as
well as the functioning of our own body, brain and behaviour. As a result,
science has pervaded every aspect of our existence while the number of
scientific disciplines and the knowledge they represent have exploded (Zwart
2010). The rapidly growing amount of information has made communication
between disciplines increasingly difficult. Consequently, there is an urgent need
for integrating theory that can assist in reconciling scientific knowledge (Gilbert
1983).
The search for theories offering integrating frameworks can be regarded as
‘centripetal’ science. The search for centripetal concepts is not a new
phenomenon in science. There are quite a few examples of early attempts to
bring order into the chaos of phenomena observed in the world. A well-known
example is the Scala Naturae (natural ladder) in which the Greek philosopher
and naturalist Aristotle ranked natural phenomena by decreasing perfection,
from spiritual and divine beings to man, animals, plants and finally rocks and
formless matter. The Scala Naturae inspired the medieval ‘Great Chain of
Being’, at the top of which Christians placed angels and God. More recent
examples of centripetal approaches are Mendeleev’s periodic table of the
elements and the tree of life that represents an evolutionary viewpoint on the
formation of species. Mendeleev’s periodic table of the elements logically
arranged atoms and their reactive properties and allowed new elements to be
predicted where gaps occurred in the scheme. Darwin’s Theory of Evolution
offered a logical mechanism for the processes that cause global biodiversity.
Both Mendeleev’s and Darwin’s theories have unified the separate domains of
atoms and organisms, respectively. However, an overarching theory that
connects domains as different as fundamental particles, atoms, molecules, cellbased organisms and technical ‘organisms’ is still lacking. From the point of view
of scientific integration, the absence of such an overarching theory must be
considered a gap at the heart of science.
To fill this gap, a theory is needed that explains the hierarchical emergence of
increasingly complex system types. The suggestion that system hierarchy
represents a knowledge gap may seem strange because the subject has been
discussed in literature for decades and there appears to be a fair degree of
General introduction
11
consensus about how to rank system types from the small to the large. A typical
example of a ranking in the literature is the ‘ecological hierarchy’. This hierarchy
includes quarks, hadrons, atom nuclei, atoms, molecules, organelles, cells,
organs, organisms, populations, colonies, communities, ecosystems, planets,
solar systems, galaxies, clusters of galaxies and the universe. The simplicity of
this sequence is alluring but camouflages many serious problems concerning
both hierarchy rules and the selection of the elements. What makes the problem
more complicated is that the above kind of hierarchy is used in different forms as
the basis for other studies and this leads to inconsistent reasoning and, thus,
sloppy science.
The problem of how to create a strict ranking may be solved by applying
evolutionary principles. In evolution, certain things are constructed first and other
things later. It is questionable, however, whether invoking evolution will help
because Gould’s replayed tape metaphor (Gould, 1989) suggests that every
hypothetical re-run of the evolutionary process must produce a different
outcome. Consequently, it is highly uncertain whether a particular species could
emerge twice. Using the same reasoning and broadening it to all systems, it can
be stated that every re-run of the universal evolutionary tape could result in a
differently organised universe. The idea that re-running the evolutionary tape
results in unpredictable new situations seems to be inspired by chaos theory.
Chaos theory states that certain dynamic processes cannot be predicted in
advance because the nonlinear development of their future states critically
depends on initial conditions. Lorenz popularised this critical dependence as the
butterfly-effect: a wing beat of a butterfly on one continent may create a
thunderstorm on another.
If the butterfly-effect were the only rule shaping the universe’s evolution, the
process would be utterly unpredictable. But, if it is true, how can we explain hat
spectral analyses of remote stars show that these stars harbour the same
physical particles, e.g., atoms and molecules, as the earth? And how can we
explain that the cell holds such a central position as the basal building block of
all organisms on Earth? The general existence of atoms, molecules and cells -at least on earth -- is intriguing because it hints at the existence of a fundamental
and universal structuring process.
The major question of this thesis is whether it is possible to find rules guiding the
above mentioned universal structuring process. To answer this question, the
work focuses on circular self-organisation processes because such processes
clearly stand out from their environment as individual entities. A cycle of
interactions in a system is also regarded as a ‘closure’. Of special interest for
this thesis are closures based on thermodynamic-driven self-organisation. A
system based on this type of closure may show a next closure and in this way
produce the next closed system. By repeating this process a long sequence of
12
Chapter 1
closure-based systems can be built. Using a specific type of closure, a long
series of closure steps can now be recognised in the emergence of increasingly
complex particles and organisms in the universe. The proton, the atom, the
molecule, the cell (prokaryotic and eukaryotic), the multicellular organism and
the animal are examples of increasingly complex closed systems. Every one of
these systems can be considered an individual entity showing its own system
dynamics in the environment in which it ‘operates’ (i.e., functions). For this
reason, these systems are labelled ‘operators’ in this thesis.
The search for hierarchic order described in this thesis originated in an
ecotoxicology study aimed at integrating toxicant effects on terrestrial organisms
and ecosystem functioning. To properly integrate these effects, all targets for
change in ecosystems needed to be arranged strictly and hierarchically. This
action initiated the quest for a strict hierarchy of levels of complexity in nature. In
turn, this quest for levels led to a search for a general particle hierarchy, as this
was considered the backbone for all other hierarchies. As is described in the
next paragraph, the major ideas that were developed during this ecotoxicology
study became the starting point of the present theory.
As an introduction to particle hierarchy, this thesis will present an overview of the
field of system studies and the role of particle hierarchy therein. The introduction
is divided into two parts: a summary of system science and a synopsis of
particle-related discoveries in physics and biology. After this general
introduction, the work of Teilhard de Chardin (1881-1955) is discussed. Teilhard
de Chardin was one of the first scientists to develop the idea of a general
particle hierarchy that included physical particles and organisms. The
introduction is followed by a process-based analysis of all the individual closure
steps in the operator hierarchy. Finally, two topics are discussed that are of
general importance to this study’s context. The first topic examines whether the
operator hierarchy can be regarded as a falsifiable methodology. The second
discusses thermodynamics as a necessary basis for all systems. The
introduction ends with an outline of the subjects discussed in chapters 2 to 9 of
this thesis.
General introduction
13
Problems with hierarchy
The realisation that quite a few things were wrong with existing system
hierarchies came in 1992. In that year I was involved in an ecotoxicology study
aimed at integrating the findings of roughly two dozen PhD projects dealing with
separate aspects of the Netherlands Integrated Soil Research Programme
(NISRP). The PhD candidates had investigated different aspects of soil
ecotoxicology. Their topics ranged from the availability of toxicants in the soil,
through absorption-desorption balances between soil and the surface of roots,
cuticle or skin, to toxicant uptake with food and the complex dynamic balance of
the internal concentration in organisms, followed by the effects on targets in
organisms and the impacts on individuals and their functioning in ecosystems.
The integration study covered many levels of ecological organisation and
needed a framework for ecological hierarchy that could act as a backbone for
the integration process.
Figure 1.1. A system hierarchy
that rests on the intuition of lower
level elements taking part in the
organisation of higher level
elements
A literature search produced a fascinating plethora of ecological hierarchies,
most of which were variations on the following themes: atom, molecule,
organelle, cell, organ, organism, population, community, ecosystem, planet, etc
14
Chapter 1
(e.g., Miller, 1978; Koestler, 1978; Sheehan, 1984; Naveh and Lieberman, 1994;
Laszlo, 1996; Newman and Jagoe, 1996; Nederbragt, 1997; Hœgh-Jensen,
1998; Korn, 2002) (Fig. 1.1). The logic behind these rankings seemed to rest on
the intuitive notion that lower level objects are ‘taking part’ in the organisation of
higher level objects; that is, atoms take part in the organisation of molecules,
which take part in organelles, which take part in cells, and so forth.
Upon closer look, however, the latter examples of rankings were generally found
to be inconsistent. To begin with, their hierarchy seemed in many aspects to be
constructed specifically from the point of view of multicellular organisms because
unicellular organisms do not consist of cells and do not have organs. In fact,
some hierarchies seemed to ignore that organisms may differ in complexity,
from bacteria and eukaryotic cells, through bacterial and eukaryotic
multicellulars to animals. Another interesting observation was that the
hierarchies generally could not be considered as representing a historical
sequence. For example, while atoms integrate to create molecules, molecules
do not first integrate to create organelles, which then integrate to form cells. The
reason for this is that as long as nature has not constructed a cell, neither the
context nor the means exist for organelle evolution. The same reasoning applies
to organs in multicellular organisms. Furthermore, it may seem confusing that
the ranking was based on two different types of transitions. The first transition
type involved particles creating next particles, for example, the step from the
atom to the molecule and from the unicellular organism to the multicellular
organism. The second transition type involved the interaction of particles in
systems that were not particles themselves. With respect to the latter transition it
is suggested that organisms take part in populations, which take part in
communities, which take part in ecosystems. Yet it is hard to imagine how
organisms are first integrated into populations, then into communities and then
into the ecosystem. As Schultz (1967) has indicated, a community, for example,
is not a subsystem but a conceptual part of an ecosystem. Consequently,
organisms directly contribute to the ecosystem while a focus on one specific
interaction can be used to group the organisms without such grouping
representing a physical object. For example, focusing on reproduction leads to
recognising clones or populations, focusing on feeding relationships leads to
food chains, focusing on engineering relationships leads to a ‘who-uses-theconstruction-of-who’ chain and focusing on interactions between organisms of
different species leads to a community viewpoint.
Based on the above considerations, it was decided that the generally used
ecological hierarchies did not offer a strict enough basis for the integration study
because their logic needed to be improved. Consequently, a stricter analysis of
ecological hierarchy had to be developed. To keep things as simple as possible,
the individual organism was used as the foundation of the new hierarchy. A
focus on individuals would allow the association of toxicant effects with:
General introduction



15
interactions between individuals in ecosystems (e.g., populations, food
chains)
complexity levels of individuals (e.g., bacteria, protozoa, plants and animals)
targets in the internal organisation of individuals (e.g., atoms, molecules,
organelles, cells, organs)
These three aspects can be regarded as separate ‘directions’ for discussing
hierarchy. If hierarchy is discussed in one of these directions, then the aspects
of other directions cannot be included without developing mixed reasoning and
an inconsistent hierarchy. For example, transitions from bacteria to eukaryotes,
and from eukaryotes to multicellulars, can be discussed, but this ranking cannot
be continued with populations as the next step because multicellulars are not the
only organisms that form populations. Prokaryote cells, eukaryote cells and
multicellulars also form populations. Similarly, prokaryote cells have an internal
organisation, eukaryote cells have an internal organisation and multicellulars
have an internal organisation. Because all three directions for analysing
hierarchy are accessible from every organism, these directions involve
independent hierarchies and should not be mixed. Moreover, because of their
independence, these directions can be considered separate dimensions for
analysing hierarchy (Fig. 1.2).
The search for a strict hierarchy in ecotoxicology started with the ranking of
organisms of different complexity levels, from bacteria to eukaryote unicellulars
to multicellular organisms. These levels were subsequently applied when
analysing the internal organisation of multicellular organisms because the cell
level is again found in a multicellular organism. To further extend this logic to the
cell’s interior, the levels of abiotic particles also had to be included because the
cell’s interior harbours many elements smaller than cells, such as molecules,
atoms, etc.
16
Chapter 1
Figure 1.2. Hierarchy dimensions in nature’s organisation: (1) hierarchy of how
organisms form complex systems of interacting particles and/or organisms, (2) hierarchy
of how organisms self-organise to form higher level physical particles and organisms, (3)
hierarchy of the internal organisation of organisms. The same logic can be extended to
physical particles as well as multicellular organisms.
It was decided to analyse the complexity of organisms as if this complexity
resulted from a range of construction steps, for example, from prokaryotic to
eukaryotic and from unicellular to multicellular. After every step, a new system
emerged that showed new properties typical for its complexity level. Every new
property represented a new target for stressors. Consequently, the ranking of
new organism properties also offered a ranking of targets in an organism. The
following example illustrates this target ranking.
When looking at a unicellular life form, such as a bacterium, its membrane and
internal autocatalytic chemistry are the main system properties. These
properties simultaneously represent the major targets for stressors. New targets
arise as soon as a bacterium gains internal membranes and becomes
eukaryotic. The nuclear membrane, the processes in both compartments, and
the exchange of information between nucleus and cytoplasm form new targets
for stress. Going from eukaryotic cells to eukaryotic multicellulars, new stressor
targets develop in the structural binding of the cells and in the chemical
communication between the cells that make it a multicellular life form.
While analysing complexity levels of stressor targets in organisms, the question
emerged whether the ranking method that was developed could be applied to
non-living systems as well. In other words, would it, for example, be possible to
distinguish between atoms and molecules in the same way as between
unicellular and multicellular organisms? If this were possible, a general
General introduction
17
framework for complexity levels of both organisms and particles could be
created that would implicitly offer a rationale for organising all targets for toxicant
action. This promising vista was the reason that the major subject of study
turned towards developing a uniform method to rank all complexity levels
including both the non-living particles (e.g., atoms, molecules) and the
organisms (e.g., bacteria/archaea, eukaryotes, multicellulars).
Before examining the construction of a strict ranking for the complexity of abiotic
and biotic particles, the next paragraph presents an overview of hierarchy
theoretic ideas and the discovery of complexity levels of particles and
organisms. This information establishes a context for the general ‘particle’hierarchy developed in this study.
Hierarchy theory: a selection of major ideas
In the mid-20th century, Von Bertalanffy (1950, 1968) envisioned a kind of
general systems theory that would one day connect a broad range of scientific
disciplines. A strict approach to hierarchy can be regarded as a prerequisite for
any general systems theory. An early review by Feibleman (1954) shows that
many general hierarchy rules had already been identified in the 1950s, including:
 Each level organises the level or levels below it plus one emergent quality.
 Complexity of levels increases upward.
 In any organisation the higher level depends on the lower.
 For an organisation at any given level, its mechanism lies at the level below
and its purpose at the level above.
 Time required for a change in organisation shortens as we ascend the levels.
 The higher the level, the smaller its population of instances.
 It is impossible to reduce the higher level to the lower.
But what exactly is hierarchy? Hierarchy is generally regarded as a relationship
that follows two simple rules: transitivity and irreflexivity (Bunge, 1969; Simon,
1973; Webster, 1979):
 Transitivity implies that as long as A has a lower hierarchical position than B,
which, in turn, has a lower position than C, the hierarchical position of A is
automatically lower than C.
 Irreflexivity means that A can never hold a hierarchical position below or
above itself.
From these rules antisymmetry follows because if we suppose a symmetrical
relationship in which A is lower than B, and B lower than A, then, by transitivity,
A would be lower than A, which is forbidden because of irreflexivity.
Examples of hierarchical relations are ‘is larger than’, ‘is older than’ and ‘is a part
of’.
18
Chapter 1
The simplicity of the above definition does not mean that applying hierarchy to
real world systems is always straightforward. First of all, problems arise if levels
depend on different hierarchy rules, resulting in what can be considered a mixed
hierarchy. Secondly, the demand of transitivity (if A is hierarchically below B and
B is below C, then A is below C) creates a hierarchy, but does not forbid
inserting an element into a hierarchy or taking an element from it. For example,
for any situation where a system A is hierarchically lower than B, a system A*
may be inserted into the hierarchy as long as A* is hierarchically above A and
below B. Yet, such ‘flexibility’ or ‘generality’ precludes constructing a strict
hierarchy in which the elements remain at fixed levels. To solve this problem, the
elements must show a non-transitive hierarchical relationship. This implies that
when A is hierarchically below B, and B is hierarchically below C than A is not
hierarchically below C. The difference between a transitive and non-transitive
relationship is comparable to the difference between using ancestry or
parenthood as a ranking rule. Using ancestry, both the grandparent C and the
parent B are ancestors of A. Using parenthood (and excluding the possibility of
incest), the situation is different because when A is child of B, and B is child of
C, then A cannot be a child of C.
Finally, there is the question of how to deal hierarchically with a circular
relationship? In fact, a circular relationship no longer represents a hierarchy
because it has lost transitivity, non-transitivity and irreflexivity. Consequently,
there is no way of determining which interaction is first or last. Heylighen (2010)
used this aspect to indicate that a circular system must also be regarded as a
new unit for defining time. Therefore, when circularity occurs, a refocusing of
hierarchy analysis is implied. Before circularity, the analysis of hierarchy focused
on relationships between the elements that create the circularity. When the
circularity is present, a new focus is required that ranks the holistic properties of
the circularity in comparison to the holistic properties of other circularities.
The above peculiarities may well be the reason that a generally accepted
analysis for hierarchy in natural systems has not been found. Yet, Von
Bertalanffy (1968) has indicated that when it is found, a ‘…general theory of
hierarchic order obviously will be a mainstay of general systems theory’.
Not everyone considers the subject of hierarchy equally important (e.g.,
Guttman, 1976). One reason for this may be that the early theories suffer from
an overly general attitude. They consider everything a system that is based on
lower level elements and that is affected by its environment. Given this
viewpoint, the theory becomes so abstract that all contact with specific real world
examples is lost. Another reason may be that general system theorists have
sometimes considered one-dimensional representations a sufficient basis for
generalisation. An instructive example is the idea that energy is the aspect that
rules an ecosystem’s organisation (e.g., Odun’s systems ecology, 1994).
General introduction
19
Despite the advantage of simplicity, any low-dimensional representation always
neglects aspects. For example, an energy-based approach to the ecosystem
neglects the structural effects of ecosystem engineers, the genetic effects of
selection and the role of dispersal, which according to the DICE approach
(Dispersal, Information, Construction and Energy, see Jagers op Akkerhuis,
2008) are the required inputs for minimal models of ecosystems.
Another reason why everyone does not consider hierarchy equally important
may be that the reductionism-versus-holism debate has troubled the objective
hierarchical analysis of systems. This debate started with Aristotle, who in his
Metaphysics (Book, H, 1045: 8-10), coined the phrase: "The whole is something
over and above its parts, and not just the sum of them all".
Reductionists generally disagree with Aristotle’s statement because they
consider that a system with all its properties, including emergent properties, can
be completely explained as the result of the interactions between the constituent
parts. According to their viewpoint, a thorough understanding of a system’s
parts, their properties and interactions allows a full understanding of a system’s
functioning, at least in principle. Consequently, a system cannot be more than its
interacting parts. This reductionist viewpoint has led to a search in systems for
constituting parts and their functional properties, the parts of the parts and their
properties, etc. For this reason reductionist theory shows a close link with a
hierarchical analysis of a system’s construction. Examples of hierarchical
approaches are the ‘worlds within worlds’ approach and abstract simplifications
of it represented by Chinese boxes or Russian babushkas (e.g., Simon, 1962;
Koestler, 1967; Lazslo, 1972, 1996).
In comparison to the reductionist viewpoint, a holistic viewpoint strongly
emphasizes the system’s emergent properties and the way these affect the
functioning of its parts. Illustrative examples of how a whole system affects its
parts are the presence of genes and organelles in the cell, the presence of
castes in social insect colonies, and the presence of organs and the brain in
multicellular animals. These and similar phenomena are fully explained by
physical laws (reductionist point of view) while at the same time their typical
realisation cannot be explained in a purely ‘bottom up’ way; they can only
develop in an environment where the higher level of organisation (cell, colony,
animal) is exposed to selection (e.g. Turchin 1977, Witting 1997, Wilson &
Wilson 2007). In addition, the holists stress that the cooperating parts may show
emergent properties that, in their typical realisation, are extremely hard to
predict. For example, the thought of a muffin may give a person a hungry feeling
and even make his mouth water. However, it is virtually impossible to predict
deterministically, from the level of the interacting nerves, why that person
thought of a muffin and not a slice of ham. What makes predicting a thought
even more difficult is that the thought may be guided by a continuous flow of
20
Chapter 1
internal and external causes, such as reading a story in a cooking magazine and
the way the reader’s eyes process the magazine’s perception.
Another aspect of emergent properties that reductionists sometimes question is
the acclaimed discontinuity between two system states at the moment of
emergence. The idea is that emergence can be regarded as a gradual trajectory
(in time and/or space) along which it is not possible to say exactly when an
emergent property is present. This idea is sometimes used to argue that it is
useless to focus on emergent properties in relation to the analysis of
organisational transitions. Indeed, if we look closely at the moment of its
occurrence, the presence of an emergent property, such as a closure, may not
be easily recognizable; during the transition phase, a closure may switch
between being present and being lost. This switching can be compared with a
person who walks between two rooms. When going through the door, the person
is not entirely in the first room and not entirely in the second room. Yet, such
presence/absence dynamics during a transition state offer no arguments for
denying a transition. At a certain point a person has fully passed the door and
has entered the second room. What may trouble this discussion is that in nonclosed self-organised systems, such as a tornado, the transition is indeed
absent because the emergent property, the whirl, is based on a gradient and
does not involve a closed topology acting as a system limit. In the case of a
tornado, there is no ‘door’ to walk through; there are only two parts of the same
room directly connected by a long winding hallway.
The above discussion shows that both the deterministic and the holistic
viewpoints have their virtues and that they supplement each other when it
comes to understanding the functioning of a system.
In addition to holism and determinism, another relevant viewpoint for particle
hierarchy is the evolutionary viewpoint. The evolutionary viewpoint focuses on
the way in which small particles create more complex particles, which, in turn,
create even more complex particles, etc. in a long hierarchy of particle
complexity. Because this viewpoint considers the construction of increasingly
complex particles, it would be logical to call it constructivism. The concept of
constructivism, however, already has various scientific uses two of which are
particularly relevant to particle hierarchy. One use relates to constructivist
epistemology. Constructivist epistemology is an epistemological perspective in
philosophy asserting that scientific knowledge is constructed by scientists and
not discovered from the world. The other use relates to a constructivist viewpoint
in mathematics that originated in a 1938 paper by Gödel (Gödel, 1938). In this
paper, Gödel states that only those mathematical objects that can be
constructed from certain primitive objects in a finite number of steps can exist.
Gödel’s mathematical constructivism offers a useful analogy to real world
General introduction
21
systems, if it is assumed that reality emerges as a large hierarchy from primitive
elements constructing more complex elements.
Whether it is wise to call it a constructivist viewpoint or not, the emergence of
higher level particles on the basis of lower level ones is considered in this thesis
the most logical perspective for analysing particle evolution. Two general
observations support this assumption. The first is that scientific measurements
solidly support that the universe shows a long history of more or less globular
expansion. When looking back in time, any globular expansion implies some sort
of origin at its centre. The second is that the concentration in a small origin of all
matter in the universe implies that material is heated. Heating generally
decomposes complex structures to lower level elements. The combination of
these two points strongly suggests that the universe once started as a small
volume with a high energy density (high temperature) in which only low
complexity particles existed. The latter shows that evolutionary particle hierarchy
is a likely candidate for a major organising principle in science. The following
historical overview presents a selection of authors that have substantially
contributed to a hierarchical construction theory of particles.
Whitehead (1929) presents a system of process metaphysics that defines the
world as a process of becoming of actual entities. These entities are both
process and outcome because they are both the subject (as the process of
becoming) and the superject (as the result of the process). Whitehead reasons
that processes are the basis of all entities: ‘… we are faced with the question as
to whether there are not primary organisms which are incapable of further
analysis. It seems very unlikely that there should be any infinite regress in
nature. Accordingly, a theory of science which discards materialism must answer
the question as to the character of these primary entities. There can be only one
answer on this basis. We must start with the events as the ultimate unit of
natural occurrence.’ (Whitehead, 1925). Whitehead refers to the events
representing the ultimate units of occurrence (the lowest possible level) as
actual entities or actual occasions, while “whatever things there are in any sense
of ‘existence’, are derived by abstraction from actual occasions” (Whitehead,
1929). When actual occasions interact and integrate into macroscopic objects,
they form a nexus (plural: nexũs). All macroscopic things we see are nexũs: a
tree, a table, a human being, a star, etc. The composition of the actual entities
contributing to a nexus can change. Whitehead also refers to individual actual
entities as creatures because they are the product of self-creation. He uses the
concept of the organism for actual entities as well as for nexũs of actual entities.
The work of Teilhard de Chardin (1959, 1969) focuses on physical and biological
particles. He identifies particles on the basis of organisational properties that he
describes as formedness and centeredness. He ranks the identified particles in
22
Chapter 1
order of emergence. His hierarchy includes abiotic particles, single-celled and
multicellular organisms and a future holistic state for society, Point Omega.
Because his ideas about the organisation of systems are still relevant, his work
will be discussed in more detail later in this introductory chapter.
Simon (1962) focuses on the role of modularity in the evolution of complexity. He
convincingly argues that it is much more efficient for nature to work with
assembled modules than to start every assembly process from scratch. In a
quantitative example, he introduces the watchmakers Hora and Tempus,
assembling watches consisting of 100 parts. Hora and Tempus are regularly
disturbed in their work. Each time they are disturbed, they lose the unit they are
working on. Hora then starts to assemble units in pieces of 10 and then puts
these modules together in a final assembly. When disturbed, he only loses the
last 10-piece module. Tempus continues to build his watches from 100 individual
elements. Every time he is disturbed, he has to start again. Simon shows that it
is easy to calculate how unsuccessful Tempus will be and demonstrates the
advantage of modularity in the evolution of system types.
Koestler (1967, p.301) introduces the holon for the elements at the nodes of
hierarchical relations and states that ‘The holons which constitute an organismic
or social hierarchy are Janus-faced entities: facing upward, toward the apex,
they function as dependent parts of a larger whole; facing downwards, as
autonomous wholes in their own right’.
Turchin (1977) introduces the concept of the metasystem transition (MST) to
describe structural and functional aspects of evolutionary steps from one level to
the next, from system S1 to system S2, that occur when, given two or more
systems S1 (S11 to S1N) that may or may not show variations, the highest control
level of the individual systems S11 to S1N becomes itself controlled. Turchin also
considers every metasystem transition a quantum of evolution and furthermore
introduces the law of the branching growth of the penultimate level. The law
states that only after the formation of a control system S will it become possible
for the subsystems Si to multiply and differentiate.
Varela (1979) introduces the concept of autopoiesis (Greek for “self-making”) for
a system in which the parts, including a boundary, support dynamics that
recreate every aspect of the system. He defines autopoiesis as follows: “An
autopoietic machine is a machine organised (defined as a unity) as a network of
processes of production (transformation and destruction) of components which:
(i) through their interactions and transformations continuously regenerate and
realise the network of processes (relations) that produced them; and (ii)
constitute it (the machine) as a concrete unity in space in which they (the
components) exist by specifying the topological domain of its realisation as such
a network.” Including the concepts of ‘regeneration’, ‘production’ and the
General introduction
23
‘concrete unity in space’ in the definition creates a natural association with the
dissipative catalytic processes and membrane as a subset of the cell’s
properties but makes autopoiesis unfit for describing the brain.
Eigen and Schuster (1979) introduce the hypercycle theory. A hypercyclic
process is created when catalysts that can individually perform a cyclic catalytic
process interact. Their interaction results in a second-order cyclic process: a
hypercycle. Eigen and Schuster (1977, 1978a, 1978b, 1979) and Kauffman
(1993) have worked out in detail aspects of the catalytic hypercycle.
Gánti (1971, 2003a, 2003b) has developed the chemoton theory. Gánti takes the
cell as the building block of life. The chemoton is constructed as a minimum
complexity representation of the cell, and thus ‘minimal life’. The chemoton
theory is based on networks of chemical reactions in which cyclic reactions
fluidly represent the information required to produce the system. RNA or DNA
are not required at this level of simplicity. The chemoton theory is supported by
cycle stochiometry, a theoretical framework. Stochiometry refers to quantitative
calculations for describing the number of reactants and products before and
after a reaction. A chemoton consists of three compartments: (1) a chemical
motor system, (2) a chemical boundary system and (3) a chemical information
system. Chemotons have been suggested as an appropriate model for the cell
and, as such, for the origin of life. Gánti recognises absolute and potential
criteria for living organisms. The absolute criteria are (1) individuality, (2)
metabolism, (3) inherent stability, (4) the presence of a subsystem that carries
information for the system, and (5) the regulation and control of processes. The
potential criteria are (1) growth and reproduction, (2) hereditary change, and (3)
mortality. According to Gánti’s definition, any technical intelligence (e.g., an
intelligent robot) cannot be regarded as life. Moreover, too narrow a focus on
minimal life does not allow the definition of life, in general, because even the
most excellent definition of a least complex cell has little to say about the rules
that define multicellularity.
Jaros and Cloete (1987) propose a biomatrix of interacting doublets. They
regard these as comparable to Koestler’s holons. Every doublet combines an
endopole, representing its internal organisation, and an exopole, representing
interactions with the environment. As an example of three levels of interacting
doublets, Jaros and Cloete choose the cellular level, the organism level and the
societal level.
Heylighen is one of the advocates of the importance of closure in system
science. Closure has a long history. Early references to closure can be found,
for example, in Wilson (1969). Wilson points out that natural boundaries may
result from minimum interactions or ‘some form of closure, either topological or
temporal’. In its most general form, closure indicates the invariance of a set
24
Chapter 1
under an algebra of transformation. Thus, when an operation is performed on
the elements of a set, the operation’s products are still elements of the set.
Closure thus indicates “the internal invariance of a distinction (or distinction
system) defining the system” (Heylighen, 1989a). As a variation on this theme, in
cybernetics a system is said to be organisationally closed if its internal
processes produce its own organization (Heylighen 1990). A well known
example of such organisational closure is autopoiesis (Varela 1979). Heylighen
proposed relational closure as a means to deal with systems that do not show a
primitive level because every element is defined relative to other elements
(Heylighen, 1990). Circular closure relationships break a system from any preexisting temporal hierarchy; all relationships loose temporal order and must be
considered as occurring simultaneously (Heylighen 2010). Heylighen also
analyzed the role of a manager as an agent causing organisational closure in
mediator evolution (Heylighen, 2006). Heylighen (1989a, 1989b) furthermore
discusses how a broad range of closed structures can be defined by differently
combining four basic closure types: (1) transitive closure (or recursivity), (2)
cyclic closure, (3) surjective closure, and (4) inverse surjective closure.
Transitivity implies that if A is lower than B and B is lower than C, then A is lower
than C. Cyclicity implies the existence of inverse transformations. Surjectivity
implies the many-to-one relationships, while inverse-surjectivity implies one-tomany relationships. In addition, any of the relationships in the four classes can
be negated e.g., non-transitive or non-cyclic.
Alvarez de Lorenzana (in Salthe, 1993) sketches a straightforward application of
a construction sequence to describe the development of the universe. He states
that the universe (U) can be understood and explained assuming that ‘…variety
and complexity can be obtained only by processes of manipulation and
combination, operated on and by the initial, given, U’. Using this reasoning, he
describes a general hierarchical model in which ‘the elements of each new level
are made out of combinations of elements of the previous level’. Alvarez de
Lorenzana further states that “Evolution of and within U takes the form of a
constructible “metahierarchy” (Alvarez de Lorenzana and Ward, 1987) – a chain
of evolutionary cycles with, ideally, no missing links’. That a construction
sequence should not have missing links was recognised earlier by Guttman
(1976). Although de Lorenzana’s approach leads to a strict series of construction
steps (no missing links), the requirement that interactions between preceding
level elements have to form the next level is too rigid to accept, for example, the
eukaryote nucleus or the brain as a next level.
Maynard Smith and Szathmáry (1995, 1999) propose another constructivist
example in the sense of lower level systems constructing higher level systems.
They focus on major evolutionary transitions, including the following examples:
A. From replicating molecules to populations of molecules in protocells
B. From independent replicators to chromosomes
General introduction
25
C.
D.
E.
F.
G.
From RNA as gene-and-enzyme to DNA genes and protein enzymes
From bacterial cells (prokaryote) to cells with nuclei and organelles
From asexual clones to sexual populations
From single-celled organisms to animals, plants and fungi
From solitary individuals to colonies with non-reproductive castes (e.g., ants,
bees and termites)
H. From primate societies to human societies and language
The rules for identifying and ranking the latter major evolutionary transitions still
need to be discussed. This need becomes clear if we analyse the points A to H
with help of the three dimensions for hierarchy that were introduced in the first
paragraph of this introduction (interactive complexity, individual complexity and
internal complexity). These three dimensions show that the ranking of the major
evolutionary transitions is based on different types of hierarchy. For example,
the transition from RNA to DNA in cells relates to the internal hierarchy, the
transition from unicellular to multicellular relates to the formation of a new
operator, and the transition from asexual to sexual and from individuals to
colonies relate to interactions in ecosystems.
So far, the above and other theoretical developments have not resulted in a
detailed, strict, evolutionary particle theory. Going back to basics, the following
paragraphs, therefore, review historical developments that have helped to give
insight into the evolutionary construction of ‘particles’, both in physics and
biology. The aim is to offer the reader these insights as a context for discussing
a comprehensive particle hierarchy in later chapters.
Hierarchy in particle physics
Particle physics has generally followed a reductionist approach because it
explains the functioning of a higher level -- for example, molecules -- from the
properties of smaller components, which in the case of molecules are the atoms
(for historical overviews, see, for example, Close 1983; Perl 1987; Walker 2009).
This methodology has strongly increased our understanding of the construction
of matter and has resulted in a long sequence of increasingly small particles,
down to the level of the quarks.
Although particles have shown a disquieting tendency not to be fundamental, a
consistent application of the reductionist approach may eventually lead to the
discovery of the smallest, truly fundamental kinds of particles, i.e., ‘the building
blocks of the cosmos’. If truly fundamental particles exist and a single logical
theory can describe the laws determining their properties and interactions,
science will then have found a ‘Theory of Everything’ (TOE) or ‘Grand Unifying
Theory’ (GUT). These labels were originally used as ironic connotations of
overgeneralised theories, but they have gained some acceptance in writing
26
Chapter 1
about particle physics (e.g., Wilczeck, 1998). As Laszlo (1994) has emphasised,
a little prudence with naming such theories may be wise because the
mechanisms described by the TOE or GUT do not reach further than particle
physics and, for this reason, will fail to explain the emergence and functioning of
more complex entities, for example, atoms, molecules and organisms. ‘t Hooft
(1992) has also pointed this out when he stated that ‘We are not at all capable of
deducing the properties of a saw-bug from the standard model, and this will
never happen, too’.
The history of particle physics dates back to the Greek philosopher Democritus
(585 BC), who proposed that every kind of matter is constructed of small,
indivisible particles, or atoms (‘that which cannot be cut’ is called ατομοσ in
Greek). According to Democritus, many types of atoms exist and they make up
all forms of matter. This atomist point of view goes back to Leucippus,
Democritus’s master. In 1808, the ideas of Leucippus and Democritus were
revived by the English chemist, John Dalton. Dalton proposed a theory in which
different chemicals were assumed to consist of indivisible particles, the atoms.
The idea of the indivisibility of atoms lasted until 1902 when Rutherford and
Soddy discovered that some large atoms could disintegrate spontaneously into
smaller atoms. About this time, Pierre Curie and his wife Marie Sklodowska
discovered that the disintegration of uranium also formed new, smaller atoms,
namely radium and polonium. In both cases, the disintegration of large atoms
into smaller ones indicated that atoms were not indivisible. Yet, the divisibility of
atoms did not immediately give further information about a precise substructure.
A few years later, in 1909, Ernest Marsden carried out experiments in which he
bombarded a thin layer of gold with alpha particles (consisting, as we now know,
of two protons and two neutrons) (Geiger and Marsden, 1913). His finding that
only 1 particle in 10000 bounced back suggested that atoms mostly consisted of
empty space, and that their mass was concentrated in a minute sub-volume. In
1911, Marsden’s work and other experiments led Rutherford to conclude that the
positive charge and the mass of atoms were located at the centre of an atom in
a compact nucleus orbited at a large distance by electrons.
At the same time several factors indicated that nuclei were composed of still
smaller elements. One was the regularity of Mendeleev’s periodic table of the
elements (1896). Another was Rutherford and Chadwick’s observation that the
bombardment of nitrogen nuclei with alpha particles formed hydrogen nuclei. In
1932, Chadwick discovered the neutron, and in 1935, Yukawa postulated that
the proton(s) and neutron(s) in the nucleus were held together by the exchange
of a small particle, which he called a pion. This particle was discovered in 1947
by Powel (Occhialini and Powel, 1947).
General introduction
27
Later, it was demonstrated that every pion consists of two minute particles, and
protons and neutrons consist of three. Gell-Mann named these particles quarks.
Quarks were never found alone but always in pairs or triplets. That quarks were
never found alone was explained by a theoretical model. The model states that
the force that holds quarks together becomes stronger when the distance
between the quarks increases. When the distance becomes large enough,
sufficient energy is accumulated to form a pair of new quarks in the middle.
Accordingly, separating a pair of quarks always results in two pairs of quarks.
The grouping of quarks in pairs or triplets was explained by assuming that every
quark carries a kind of charge, called its ‘color’, and that nature allows only
neutral combinations of color charge, represented either by a color and an anticolor or by the combination of three different colors.
Current theoretical studies try to describe quarks, leptons and bosons as special
configurations of hypothetical objects called ‘superstrings’. Alternative,
descriptions of fundamental particles explore dressed or undressed
manifestations of bare quarks (Greben 2010) or relationships with bifurcation
dynamics of self-organising dissipating force fields (Manasson, 2008). So far, no
evidence has been found for particles smaller than quarks.
Hierarchy in the construction of organisms
One of the most exciting events in early biology was the discovery of the cell
(e.g., Wolpert, 1995; Mazzarello, 1999). Until Antony van Leeuwenhoek (16321723) had invented the microscope, the cellular world was closed to
investigation. With access to this new tool, van Leeuwenhoek made many
observations that revealed the wonderful diversity of the small creatures that he
described as ‘animalcules’. His discoveries included many unicellular organisms,
such as protozoa in water drops, his own semen and bacteria in dental plaque.
In 1665, Hooke described structures in cork as ‘cellular’ and concluded that
these were channels for conducting fluid. However, he did not suggest that such
cells could be the basal units of life. It took many more observations and some
200 years before Schleiden (1838) and Schwann (1839) coined the idea that all
life was based on cells as the basal modules. Schleiden (1838) wrote that ‘Each
cell leads a double life; an independent one, pertaining to its own development
alone; and another incidental, as far as it has become an integral part of a plant’.
Compared to plant cells, animal cells have much thinner membranes. This was
probably the reason why it took until 1895 before Overton showed that the
animal tissue also consists of cells with a proper membrane.
Around this time, scientists also started to investigate the cell’s interior.
Important cellular discoveries in this respect were those of the internal ’vesicles’
28
Chapter 1
represented by the chloroplasts and mitochondria. Because they are large and
green, chloroplasts in plant cells were observed by early microscopists, including
Anthony van Leeuwenhoek. In 1682, Nehemiah Grew described green
precipitates in leaves. His description is considered the first report of
chloroplasts. Although the earliest descriptions of mitochondria date back to
1840, Richard Altmann was the first to specifically address these organelles,
calling them bioblasts. When studying spermatogenesis in 1898, Carl-Benda
introduced the name mitochondrion from the Greek words mitos, which means
thread, and chondros, which means granule.
Meanwhile, Louis Pasteur (1822-1895) had shown that a culture medium
remained sterile after heating. Pasteur’s observations indicated two things.
Firstly, a sterile culture medium had to be exposed to air currents before
organisms present on dust particles could colonise the medium and propagate.
Secondly, the chance that the chemicals in a given culture medium would
spontaneously combine to form living cells (’generatio spontanae’) was
apparently small or non-existent. In relation to these types of experiments,
Virchow (1855) stated that all cellular life is derived from parental cellular life
(‘omnis cellula e cellula’).
The cell became a basic concept in biology, and in 1892, Wilson (Wilson, 1892)
wrote that ‘no other biological generalisation, save only the theory of organic
evolution, has brought so many new points of view or has accomplished more
for the unification of knowledge’.
However important the cell had become, the idea that all cells originated from
cells did not last. Regardless of how small the chance, when thinking about the
emergence of life on earth, two possible explanations arise: one is forced either
to accept a chemical origin for life on earth or to speculate that extraterrestrial
life colonised the planet. Since the latter speculation contains the question of
how these earth-colonising organisms came to be, both lines of explanation
imply that a solution is required for the emergence of living cells from a chemical
environment.
A range of theories has been proposed that have helped to understand the
conditions that may be required for the emergence of life and to understand
what constructions may be typical for the first cell (Oparin, 1957;
Wächtershäuser, 2000; Eigen, 1977; Kauffman, 1993). In order not to invoke
Dennett’s ‘skyhooks’ (Dennett, 1995), his metaphor for the impossibility of top
down constructions, scaffolding is required by which simple elements can pull
themselves up by their bootstraps while creating the cell’s complex ensemble,
including the maintenance of its individuality on the basis of catalytic chemistry
and a membrane (Conrad, 1982; Martin and Müller, 1998; Gánti, 1971;
Hengeveld and Fedonkin, 2007). Although no approach has yet led to a de novo
General introduction
29
construction of cells, the first cells must have had autocatalytic chemistry and a
rather simple construction, while more complex structures, such as DNA and
organelles, gradually evolved afterwards.
To define the cell’s basic properties, the artificial life community frequently uses
the concept of autopoiesis (Maturana and Varela, 1980). Interestingly, problems
have arisen because the original definition of autopoiesis refers to a network of
component production processes that continuously constitute the system as a
concrete unity in space. Yet, the cell is an open system that depends on a
continuous influx of substrates and efflux of waste. Some scholars now reason
that, for example, in the case of spider venom, bacterial excretes and intestinal
digestion fluids, the external digestion enzymes actually form part of the network
of processes reconstructing the cell and that, for this reason, the membrane of
the cell is not the limit for the autopoietic dynamics. This controversy indicates
that autopoiesis should perhaps be more strictly interpreted, namely that the
membrane should be considered a closed interface only for its mediating role in
maintaining sufficient internal concentrations of those catalytic molecules that
are directly involved in the cyclic basis of the autocatalytic process. This
restricted interpretation of autopoiesis would sharpen Maturana and Varela’s
definition; it would more clearly explain that only the placing of the basic catalytic
molecules outside the membrane would break up the cells catalytic closure and
place it outside the autopoietic domain.
Just as van Leeuwenhoek’s invention of the microscope caused a revolution in
biology because it made cells visible, the invention of the electron microscope in
1931 (Knoll and Ruska, 1932a, 1932b, Ruska, 1986) greatly enhanced our
vision of the much smaller world hidden in the cell’s interior. By 1935, an
electron microscope had been built that had twice the resolution of a light
microscope. In later years, electron microscopes would allow detailed study of
the cell’s broad range of structures, including the endoplasmatic reticulum
(named by Garnier in 1897), the Golgi apparatus (discovered by Golgi in 1898),
mitochondria (named by Benda in 1898), and spindles and the nuclear
envelope.
An exciting development with respect to mitochondria and chloroplasts was
Margulis’ hypothesis(1970, 1981) that these organelles evolved from bacteria
that had invaded host bacteria and continued to live in it as endosymbiont (see
also Schwartz and Dayhoff, 1978; Gray, 1992; Knoll, 1992). There is currently
broad support for this theory, based on evidence that the endosymbionts have
their own genetic material, produce their own cell membranes and multiply by
cell fissure. Nevertheless, questions still remain about how the symbiosis was
established. In a recent publication Blackstone (1995) has asked attention for a
units-of-evolution approach that simultaneously accounts for the gains and
losses that the interacting cell types experience when living together compared
30
Chapter 1
to when living separately in the environment. The approach also focuses on the
difficulties both partners may encounter to communicate and, to a certain extent,
to control the other. Recent discussions on an anoxibiontic origin of mitochondria
can be found in Martin and Müller (1998), Doolittle (1998), Martin et al. (2001)
and Martin and Russel (2003). These publications propose that a strictly
anaerobic archaebacterial host, that initially was not capable of digesting larger
organic molecules, and a facultative anaerobic -proteobacterium, that could
digest organic molecules, may have been the first to establish endosymbiosis on
the basis of H2 and CO2 production by the -proteobacterium. The Archaean
host then metabolised the H2 and CO2 to methane. The latter theory also
assumes that the development of a karyos took place after transfer of genes for
the production of eubacterial membrane lipids to the Archaean host.
Forms of endosymbiosis have recently been discovered that are even more
complex than the mitochondria and chloroplasts in eukaryote cells. The
complexity of such cell-in-a-cell relationships can be understood by focusing on
the number of internal membrane layers. For example, all compartments in the
eukaryote cell are surrounded by a single membrane whilst mitochondria show
two internal membrane layers and chloroplast even have three. This method can
also be used to analyse other remarkable endosymbiontic relationships. For
example, in phototrophic cryptophytes, a red alga is contained in the rough
endoplasmatic reticulum of the host. This means that four additional internal
membrane layers can be found within the membrane of the endoplasmatic
reticulum of the host. The first is the outer membrane of the endosymbiontic red
algae. The second and third originate from the double membrane of the endoendosymbiontic phototrophic bacteria, the thylakoid membranes of which form
the fourth internal membrane layer. Another example is the alga Peridinium
balticum. This heterotrophic dinoflagellate contains a phototrophic
chrysoflagellate. A total of six membrane layers of the endosymbiont can be
found within the host cell. The first is the cell membrane of the endosymbiont,
and the second and third are from an extension of the nuclear envelope that
surrounds the chloroplast. The fourth and fifth are the outer and inner membrane
of the chloroplast, and the sixth is the membrane of the thylakoid vesicles.
(Bardele, 1997).
According to the evolutionary viewpoint, the organisational complexity of
organisms can be arranged in a large tree. The branching pattern of this ‘tree of
life’ generally offers no explicit information about the structural complexity of the
species involved, such as the presence of endosymbionts, a nucleus or the
development of multicellularity. However, this problem can be solved by
specifically highlighting the various complexity levels. Figure 1.3 (Alberts et al,.
1989) shows how the ‘tree of life’ can be enriched with information about
important levels of structural organisation. The figure shows the following levels:
the cell, the eukaryote cell and the multicellular organism. The figure also
General introduction
31
illustrates that the brain is regarded as an organ and not as a separate level of
organisation.
Figure 1.3. Representation of the ‘tree of life’ showing the historical speciation processes,
the occurrence of endosymbiosis and the closure levels of the first cells, the eukaryote
cells and the multicellular life forms (modified from Alberts et al., 1989).
A pioneer of particle hierarchy: Pierre Teilhard de Chardin
Teilhard de Chardin was one of the first scientists to combine a construction
viewpoint with a focus on particles to create a worldview that incorporated
physical particles as well as organisms in one large evolutionary hierarchy.
As a palaeontologist, Teilhard de Chardin studied the evolutionary processes of
organisms, including the human race. His studies in China included early work
on the Peking man, Sinanthropus pekinensis (presently classified as Homo
erectus pekinensis). Teilhard de Chardin would probably be less known if he had
not broadened his paleontological viewpoint to incorporate major aspects of the
evolution of abiotic systems. Using his ‘law’ of complexity-consciousness as a
central heuristic principle, Teilhard de Chardin created a unique all-embracing
representation of the evolution of complexity in the universe. His extrapolations
of evolution’s course beyond the present allowed him to speculate about
humanity’s future. His approach was unique because it was predominantly
based on particles, which he described as systems that were both ‘formed’ and
32
Chapter 1
‘centered’. Due to this focus on particles, he could integrate evolutionary
transitions between physical and biological particles into one large sequence. He
thought this sequence showed a direct link with the chronology of the
emergence of complexity and, therefore, with the genesis of the universe (ES
VIII, p. 36; 1949).
Teilhard de Chardin called the core of his theory ‘complexification intériorisante’,
or inward complexification, defining complexity as the result of relationships of
elements amongst themselves. Following this reasoning, he arrived at two
insights that are still relevant for particle–based evolution:
First, in the multitude of things comprising the world, an examination of their
degree of complexity enables us to distinguish and separate those which may be
called ‘true natural units’, the ones that really matter, from the accidental
pseudo-units, which are unimportant. The atom, the molecule, the cell and the
living being are true units because they are both formed and centered, whereas
a drop of water, a heap of sand, the earth, the sun, the stars in general,
whatever their multiplicity or elaborateness of their structure, seem to possess
no organisation, no ‘centricity’. However imposing their extent they are false
units, aggregates arranged more or less in order of density.
Secondly, the coefficient of complexity further enables us to establish, among
the natural units which it has helped us to ‘identify’ and isolate, a system of
classification that is no less natural and universal (ES V, p. 137; 1946).
These two points have remained relevant because they highlight some
fundamental principles for identifying particles, distinguishing them from other
systems and ranking them in a general hierarchy. Teilhard de Chardin’s work
includes some personal points of view, of which the following are presumably
the most relevant in the context of particle hierarchy. Firstly, he assumed a full
correspondence between complexity and consciousness. For this reason, it
made no difference to him to discuss the complexity or the consciousness of an
atom. Next, he regarded evolution as ascending to consciousness and thus
assumed that evolution would reach its zenith in some sort of highest
consciousness (Teilhard de Chardin 1959). Subsequently, he predicted that all
individuals in society would evolve to an integrated organisation/consciousness,
which he called Omega.
The use of consciousness as a synonym for complexity merges a constructional
and a functional property of multicellular animals. Unless we assume that
consciousness is an esoteric phenomenon requiring no physical basis,
consciousness is a mental process that is caused by interactions between brain
cells, and, for this reason, functionally depends on a multicellular construction.
Consequently, Teilhard de Chardin’s equating complexity with consciousness
General introduction
33
implies the presence of multicellularity at all complexity levels, including, for
example, the levels of unicellular organisms and atoms. The latter example
shows that Teilhard de Chardin’s viewpoint suffers from overextending the term
consciousness, as Ulrich (1972) has also indicated
In principle, when extrapolating his particle hierarchy, Teilhard de Chardin used
a scientifically acceptable heuristic procedure. The integration of multiple
personalities into a single organism would be analogous in this context to the
integration of atoms that create molecules, or cells that create a multicellular.
However, as will be made clear in this thesis, there are serious problems with
using the human phenotype as the basis for such integration. Furthermore, it is
hard to defend Teilhard de Chardin’s suggestion that Point Omega represents
some final evolutionary unity because every historical step in the evolution of
particle hierarchy has been succeeded by a next step. As a result, the claim that
Omega suddenly would represent a final situation must be considered illogical
and hence improbable, even though, strictly speaking, the successful
occurrence of past steps does not guarantee the possibility of future steps.
Even though certain aspects of Teilhard de Chardin’s philosophy can be
improved, his ideas of inward complexification and true and false units as well as
his comprehensive hierarchical ranking, including abiotic and biotic particles, are
demonstrations of his visionary insight into the abstract nature of system
hierarchy and into the universality of the evolutionary process.
Introducing first-next possible closure
The concept of closure, in particular that of first-next possible structurofunctional closure, is essential to the operator hierarchy. The next paragraphs
introduce this concept.
What is closure? One of the oldest visualisations of closure is perhaps the
ancient symbol of the Ouroboros, the snake that swallows its own tail and by
doing so creates a closed structure: a circle. In more general terms, closure
relates to the situation in which a set is closed for the performance of an
operation on its elements. This leaves room for different closure types. For
example, the set of integers {....-2, -1, 0, 1, 2, 3,...} is said to be closed under the
operation of subtraction because the outcome of every subtraction is a member
of the set. The set of natural numbers {1, 2, 3,…}, however, is not closed for
subtraction because 2 minus 7 is -5, which is not a natural number.
The application of closure to system dynamics plays a major role in Heylighen’s
work (e.g., Heylighen, 1989a, 1989b, 1990). Closure adds a special property to
a system because the state space of a closed system has become invariant (i.e.,
it does not change) under the internal dynamics (e.g., Heylighen, 1990).
34
Chapter 1
Moreover, closed dynamics allow one to distinguish between the system and its
surrounding world.
In the context of the operator theory, closure refers both to the closing process
and to the closed (dynamic) state that results from it. Consequently, a closure is
intimately linked to underlying mechanisms through the functionalities of the
elements causing the closure. Once present, the closed system with its
emergent dynamics gains mechanistic importance because it allows for
discussion about the system’s accountability as a unity, about novel dimensions
for entropy production and about the sensitivity to novel selective forces. If there
is no closure, there is no new unit that can simultaneously be the object and the
subject in interactions.
In line with other publications (e.g., Heylighen, 1995), the operator theory
distinguishes between functional and structural closure. Functional closure is
defined as a group of elements that show a closed cycle of interactions that do
not involve a physical mediating layer. Examples of functional closure are the
autocatalytic set, the pion exchange between protons and neutrons in the atom
nucleus, and the plasma strands connecting the cells of multicellulars. In
contrast, the definition of structural closure demands that interactions create a
physically closed topology. In the operator hierarchy, the focus is on a special
form of structural closure, namely the structural closure that creates a physical
boundary mediating a contained, first-next possible, hypercyclic process.
Specific examples of structural closure in the operator hierarchy are the electron
shell around the atom nucleus, the cell membrane of a single cell, and the
connected cell membranes of multicellular organisms. Structural closure is
important because the presence of a physical mediating layer is essential to
defining a system as a ‘unit’ or as an ‘individual’. For example, without the
structural closure of a cell membrane, the catalytic molecules of an autocatalytic
set would float freely in their chemical soup. They would drift apart, dilute or mix
with other sets. As long as they can mix, it is impossible to identify the catalytic
molecules as members of any autocatalytic set in particular.
While constructing the operator hierarchy, I aimed at a strict ranking of all
elements involved because only a strict hierarchy can be used as a reference or
‘yardstick’ for complexity. Here strictness indicates situations in which it is
impossible to include an element in a hierarchy or exclude an element from it
without disrupting the ranking’s logic. A consequence of a strict construction
sequence is that all elements build directly upon each other (Fig. 1.4). Thus, all
system types in a strict construction hierarchy can be seen as belonging to one
large family tree.
General introduction
35
Figure 1.4. This figure illustrates closures creating
a strict hierarchy of levels. The levels are only strict
as long as every closure represents the shortest
route to the next level system type.
Given the demand for strictness, the operator hierarchy has to be based on a
rule that always produces the next particle -- nothing more and nothing less. To
deal with the latter demand, the notion of closure has become the basic principle
in all publications about the operator hierarchy. As seen from the publications in
this thesis, the closure concept has been more precisely defined over the years.
In the first publication on this subject (Jagers op Akkerhuis and Van Straalen,
1999), the rules defining the transitions were described as ‘hypercycle formation’
and ‘compartmentation’. In the next paper (Jagers op Akkerhuis, 2001), the
focus was on major and minor transitions. Major transitions were described as
always relating to an emergent hypercyclic dynamic and an interface, while
minor transitions ‘reflect the emergent properties of major transitions occurring
earlier’. Still later, the aspect of first-next possible closure was introduced. In
Jagers op Akkerhuis (2007), first-next possible closure is defined as follows:
‘According to the operator viewpoint, an operator of type x creates a next
operator of type x+1 by means of a first-next possible closure. … The adjective
‘first-next possible’ refers to the demand that the closure must be the first
possibility for a new type of closure in system x+1 after the preceding closure
created the operator of type x.’
First-next possible closure always indicates a new type of structural and/or
functional closure. Using the combination of these closures as a means for
identifying complexity levels has worked well for constructing the operator
hierarchy. First of all, many steps in the operator hierarchy involve closures that
leave little room for discussion, such as the construction of the hadron, the atom,
the molecule and the cell. Concerning these closures, no relevant alternative
closure states exist that could be inserted between the atom and the molecule or
between the molecule and the cell. In addition, the use of structuro-functional
closure also offers a simple and robust rule for recognising higher level closure
steps. There is hardly a chance that the transition from unicellular to multicellular
will not be identified because the cells show recurrent physical bonds (structural
closure) and physiologically depend on each other (functional closure). The
same holds for the transition from a prokaryotic to a eukaryotic cell, where the
36
Chapter 1
endosymbionts/nucleus form internal compartments (structural closure) allowing
for recurrent interactions between the new internal levels (functional closure).
First-next possible closure is not just a structural state and the operator
hierarchy is not just a ranking of states. As stated earlier, closure stands for the
closed states as well as for the self-organisation processes causing closure.
This implies that for every transition in the operator hierarchy, the closure/closing
process is intimately linked to underlying mechanisms. However, because every
operator resides at a specific level in the operator hierarchy, it is an illusion to
expect that a single underlying mechanism could explain all these transitions.
Instead, the new properties created by the closure of an operator at level X have
to allow for the dynamics that enable the closure of the operator at level X+1.
Because every closure, in principle, is based on its proper mechanisms, an
overarching mechanism can only be recognised at a higher level of abstraction.
A good candidate for such an overarching ‘mechanism’ would be the process of
first-next possible closure.
As a practical rationale that identifies the lowest complexity boundary of every
closure type, first-next possible closure has allowed identifying the operators
without major problems.
Processes, closures and operators
This chapter discusses the functional and structural aspects of all the individual
transitions that formed all the different operators and/or intermediate closure
states that were required for their formation. The individual transitions are
discussed as part of the introductory chapter because the other thesis chapters
mainly discuss the higher-order logic of the operator hierarchy and focus on the
properties that certain types of operators have in common.
Because the operator hierarchy results from a construction process, a low level
of organisation is required as a start. At this moment, the fundamental particles
(the particles included in the standard model) represent the most fundamental
level of organisation known. For this reason, fundamental particles are used in
the following text as the starting point for analysing how closures create
operators. This choice, however, implies no a priori assumptions about whether
fundamental particles represent the lowest level in an absolute sense.
The sequence of first-next possible closures listed below includes situations in
which the emergence of a next-level operator is analysed as consisting of two
seemingly separate closures: one functional and one structural. The theoretical
‘addition’ of these closures must, in practice, be seen as functionally unifying
both closures into one emergent process. For example, the formation of a
General introduction
37
prokaryotic unicellular operator represents not just the addition of an interface to
an autocatalytic set, but implies a functional merger as well, making both
aspects fully dependent on each other.
As the following paragraphs focus on the construction of a type-hierarchy, the
examples deliberately represent the simplest representations of their kind.
Consequently, any real examples of these systems will generally involve more
complex hypercycles and/or interfaces.
Fundamental particles: a first-order cycle
Fundamental particles split off and reabsorb virtual particles, such as virtual
photons, in a process called self-interaction. Likewise, quarks show a continuous
emission and re-absorption of gluons. In a Feynman diagram, the emission and
re-absorption of a virtual photon by an electron or a gluon by a quark results in a
straight line for the electron or quark, with the virtual photon or gluon line splitting
off and curving back to it. The splitting-off and reabsorbing process can also be
considered a life cycle from the quark state before emission to the next quark
state, that is, after emission. Such a first-order life cycle is shown in Fig. 1.5.
Figure 1.5. An example of a first-order interaction
cycle based on quark-gluon exchange
Fundamental particles: a second-order cycle
At high energies, (e.g., in the extremely hot, minute, early universe) the weights
of fundamental particles are enormous, their distances minute and the forces
between them relatively small.
For this reason all fundamental particles, including quarks, are supposed to have
moved around in a relatively unbound state in a kind of ‘soup’. In this soup, the
exchange of virtual photons caused two electrons to repulse each other, while
the exchange of gluons between quarks led to mutual attraction. Attraction
based on gluon exchange can also be depicted as two first-order interaction
cycles that combine into a second-order interaction cycle (Fig. 1.6).
38
Chapter 1
A construction rule of the operator hierarchy states that the highest complexity
mechanism, which in this case is represented by gluon exchange, is the most
likely candidate for integrating the first-order cycles. The quark-gluon exchange
is complex because, unlike other force-conveying particles, gluons can split and
reunite.
Figure 1.6. A second-order interaction cycle
based on quark-gluon exchange
The emergence of the hadrons (in the cooling plasma)
The expansion of the universe cooled the quark-gluon plasma. This reduced the
energy density to below the level where quarks and gluons condensate into
small bundles of two or three quarks. This bundling is called confinement.
Responsible for quark confinement is the color force. This strong force originates
in the color charge and is responsible for quark confinement. Color charge is
conveyed by gluon exchange. As the result of the color force, single quarks do
not exist because, like the stretching of a rubber band, the color force increases
when two quarks are pulled away from each other. At a certain point, the energy
in the stretched ‘rubber band’ becomes equal to the energy required to create
new quarks. At that moment, the color force-field snaps and two new quarks are
formed at the loose ends. The separation of two quarks always results in two
pairs of bound quarks and never in two separate quarks.
The color charge comes in three colors: red, green and blue and three anticolors: anti-red, anti-green and anti-blue. Quarks carry a single color: red, green
or blue. Anti-quarks carry a single anti-color: anti-red, anti-green or anti-blue.
Gluons are color neutral because they carry a combination of a color and an
anti-color (for example, red/anti-green). Observations have shown that the color
charge of a system of interacting particles must always be neutral and that the
color charge must always be conserved in interactions between systems. It is
General introduction
39
quite special that gluons carry color charge because other force-carrying
fundamental particles, such as the photons that convey the electromagnetic
force, do not carry charge themselves.
The emission of a gluon changes a quark’s color because the gluon takes a
color/anti-color combination when emitted. Thus, when a blue quark emits a
blue/anti-red gluon, the quark changes color to become red (blue=red + [blue +
anti-red]).
The color force is considered the mechanism causing the confinement of exactly
two quarks in mesons and exactly three quarks in baryons. Mesons are
relatively unstable because a quark and an anti-quark quite easily annihilate
each other. A baryon is relatively stable. From all baryons, the proton is the most
stable because it is based on the perfect combination of three quarks with
different colors: red, green and blue.
From the perspective of the operator theory, particles consisting of two or three
quarks are called hadrons. Hadrons combine two closures: the hypercyclic
interactions resulting from second-order production-absorption cycles based on
gluons, and the closure resulting from the confining influence of the color
charge. The latter closure occurs if temperatures are low enough (Fig. 1.7).
Figure 1.7. Formation of a hadron resulting from
the combination of a second-order quark-gluon
cycle and quark confinement
Hadrons: a first-order cycle
After creating hadrons, adding more quarks (assuming this will yield a stable
system type) does not create a new system complexity. Higher complexity would
be possible, however, if a hadron could support a cyclic process that could be
used for higher order interactions. In the context of the operator theory, we
should now look for the most complex mechanism allowing the simplest, new,
40
Chapter 1
first-order cycle. The latter is represented by the emission and absorption of
small hadrons. In this process, pions, which consist of a quark and an antiquark, are the most important unit for two reasons. Firstly, pions are the lightest
and therefore 'cheapest' to produce. Secondly, pions are the most stable of all
mesons and can bridge a relatively large distance before they disintegrate. In
fact, the swarm of pions surrounding a hadron is so dense that it represents the
main part of a hadron’s mass (Fig. 1.8).
Figure 1.8. A first-order interaction cycle based on
hadron-pion exchange
Atom nuclei: a second-order cycle
From the viewpoint of the operator hierarchy, the pion exchange between
hadrons represents a second-order interaction cycle (Fig. 1.9). The exchange is
an important mechanism for two reasons. Firstly, the force of the pion exchange
is strong at small distances. It is so strong, in fact, that it overpowers the
repelling forces between positively charged protons and, for this reason, can
make protons and neutrons lump in small units, each called a nucleus.
Secondly, without pion exchange, the universe would be devoid of neutrons
because any separate neutron has a 15-minute half-life. At this decay rate, and
without new neutrons being formed, the universe would have lost virtually all its
neutrons within a few hours. By continuously exchanging pions, the protons in
the nucleus change state to become neutrons and the neutrons change to
protons. Thus, the neutrons are never ‘pure neutrons’ and cannot easily decay.
General introduction
41
Figure 1.9. A second-order interaction cycle based
on hadron-pion exchange
The electron shell
Nuclei were never alone in space. In the early universe they were surrounded by
a dense ‘soup’ of quickly moving particles, including many other multi-hadron
particles with which they could interact/mix. As long as the kinetic energies were
high enough to overcome the repelling force of the positive charge of the
protons, the ‘identity’ of the multi-hadron particles could change. The system
changed when temperatures in the universe lowered because of expansion. At
temperatures below 3000 oK, electrons became slow enough to be captured by
the positive charge of the protons. After that moment, an environment that
contained free electrons and bare nuclei condensed into the current state where
the nuclei are surrounded by a number of electrons that fits closely to the
number of protons in the nucleus. The result is a zero or close to zero net
charge and the ‘interfacing’ of the nucleus by the electron shell inhibiting
interactions between nuclei.
The atom
The atom results from combining the preceding two closures: the closure
creating the hypercyclic interaction of the nucleus and the closure creating the
interfacing by the electron shell (Fig. 1.10). These two closures together allow
atoms to emerge.
Interestingly, because the operator theory demands a second-order interaction
cycle as the hallmark of a nucleus, the system consisting of a proton and an
electron, which is generally regarded as a hydrogen atom, is not considered an
atom, but an atom-like interaction system.
42
Chapter 1
Figure 1.10. Formation of an atom resulting from
the combination of a second-order hadron-pion
cycle and an electron shell
The multi-atom (or ‘molecule’)
After the atom, the next most direct step in system complexity must come either
from a closure in the internal structure of the atom, which the laws of nature do
not allow, or from physical connections between atoms, e.g., a transition from a
single-atom state to a multi-atom state. Multi-atom systems show a recurrent
interaction between atoms due to their electron shell: they exchange electrons
between their shells in what are considered ‘covalent bonds’ (Fig. 1.11). This
exchange can result in grids (e.g., in metals) or in complex three-dimensionally
organised structures (e.g., molecules).
Figure 1.11. Formation of a multi-atom resulting
from covalent bonds between atoms
General introduction
43
The first-order catalytic reaction cycle
Adding atoms to any multi-atom system may create more complex molecules,
but these will always be multi-atomary. For a next closure type, the operator
theory suggests that one should look for the most complex emergent property.
This is offered by the three-dimensional structure of connected atoms and the
catalytic reactions which a three-dimensional structure allows for.
If we take an enzyme as an example of catalytic molecules, we can analyse its
reactions as a reaction cycle. The catalyst (C) binds to a substrate (S) and
modifies it. The result is that the substrate molecule (S) is changed into a
product (P) and the original catalyst is regained. This process represents a firstorder reaction cycle (Fig. 1.12).
Figure 1.12. A first-order interaction cycle based
on a catalyst that binds to a substrate and that is
released after the formation of a product
The second-order catalytic reaction cycle: autocatalysis
In its least complex form, a second-order reaction cycle emerges when the
product of a given catalytic reaction A equals the enzyme of cycle B and vice
versa (Fig. 1.13). Such a reaction cycle is catalytically closed, which in principle
allows it to catalyse itself as a set. This process has been named 'autocatalysis'.
Various authors, notably Eigen and Schuster (1977, 1978a, 1978b, 1979) and
Kauffman (1993) have proposed theories about catalytic hypercycles. Dittrich
and Speroni di Fenizio (2007) have named a set of self-maintaining molecules
an ‘organisation’.
44
Chapter 1
Figure 1.13. A second-order interaction cycle
based on two catalytic reaction cycles that produce
each other’s catalysts
The cell membrane
The autocatalytic set can reach a spatial limit when the molecules form a kind of
lumped aggregate or when a mediating membrane surrounds the hypercyclic
reactions. A simple lumping of molecules creates a unit, but would still allow the
mixing of such lumps and would not protect the autocatalytic set from
disturbance of reactions by molecules from outside the lump. A surrounding
layer that mediates the passage of molecules to and from the interior allows a
system that does not show such problems. This advantage may explain why
selective forces have favoured a membrane that surrounds the autocatalytic set.
From a chemical point of view, it is rather easy to form membranes from fatty
acids in pro-biotic chemical environments. Various suggestions for such
processes can be found in Oparin (1957), Claessens et al. (2007), Fanelli (2008)
and Hernández-Zapata et al. (2009), among others.
The cell
As long as the reactions of an autocatalytic set float in a larger chemical solution
and show no structural system limit, the set cannot be considered an individual
entity. A lack of individuality is problematic in recognising when an autocatalytic
set produces enough material to have made one or more copies of all its
constituents. After a full round of catalysis, an autocatalytic set has, in principle,
reproduced in some way, but what about the offspring? It -- or they -- float in
between the parent’s molecules and cannot be separated from them. As diffuse
entities, the offspring cannot be recognised as units, and they cannot go their
own way or experience individual selection.
General introduction
45
Only a system limit that acts as an interface for the autocatalytic set (which does
not imply that it acts as an interface for all atoms and molecules inside it) and
that would at least result from the catalytic process but may also participate in it,
would give the word ‘autocatalytic’ a structural meaning. Once surrounded by a
membrane, every autocatalytic reaction cycle has become part of a structural
unit, a cell, that can reproduce by forming new structural units or cells and that is
recognised as an individual because of its unique autocatalytic set and
membrane as system limit. The cell results from a merger of the closure creating
hypercyclic interactions based on catalytic molecules and the closure
surrounding the autocatalytic set with a membrane (Fig. 1.14). The two closures
together define the prokaryotic/bacterial cell.
Figure 1.14. Formation of a cell resulting from the
combination of a second-order cycle based on
molecular catalysis and a membrane
As a consequence of cell-based autocatalysis, the cell’s functioning goes hand
in hand with the degradation of environmental energy. Any system that
maintains or even increases its dynamic organisation by degrading energy from
the environment is considered a dissipative system. This does not mean that all
dissipative systems are cells. The structure of a tornado, for example, is created
by a dissipating air-pressure gradient. Yet, a tornado does not conform to the
first-next possible closures defining the cell.
Another consequence of cell-based autocatalysis is that the system must
'struggle' to stay alive because an autocatalytic set of molecules represents a
dynamic equilibrium state between degradation of the set, on the one hand, and
maintenance, growth and/or reproduction, on the other. This equilibrium shows a
minimum complexity boundary. If more molecules are lost from the set than are
required for autocatalysis, the circular chain-reaction breaks down and the
system sinks below its lowest complexity boundary. In short, it ‘dies’. To prevent
death, or in other words, to stay alive, the cell must balance the production and
46
Chapter 1
losses of the minimally required set of reaction-cycle molecules and membrane
molecules to maintain itself. If no buffers are available the minimum requirement
for survival/living is maintenance.
Cell-based autocatalysis may also lead to more than just survival. Under
favourable conditions, the production rates of catalytic molecules and other
products will exceed their degradation rates and this will increase the quantities
of all cell constituents. With a doubling of all constituents, the cell can split in two
without losing functionality. In principle, such an event marks the cell’s
reproduction.
Producing offspring represents an energy barrier. This situation is comparable to
energy barriers in enzyme reactions. Analogous to enzyme reactions, passing
the barrier represents a thermodynamically-favourable dynamic condition.
Reproduction strongly increases the chance that the gradient dissipation
processes will persist, while mortality will eventually block the process in all nonreproducing entities. When two or more cells reproduce and create dissimilar
offspring that are individually selected for their viability in a given environment,
one can start talking about Darwinian evolution.
The cell with endosymbiont/karyos
With respect to a first-next possible closure creating an internal compartment in
the cell (Fig 1.15) one may consider two possibilities: the endosymbiont and the
nucleus. As both possibilities involve structural and functional closure and
uncertainty exists with respect to the events that have led to their formation, it is
difficult to decide which of the two represents the first-next possible closure.
Support that an endosymbiont preceded the nucleus comes from genetic
analyses (Rivera, 1998) and observations of traces of endosymbionts in so
called ‘primitive eukaryotes’, the eukaryote cells that do not harbour
endosymbionts (Clark, 1995). The presence of traces suggests that the studied
cells have lost their endosymbionts in a process of evolutionary specialisation.
Additionally, recent theories about the origin of the nucleus make plausible that
genetic material coding for membrane lipids was transferred from an
endosymbiont to the host cell. This transfer allowed the host cell to create a rich
internal membrane structure that supported the development of a nuclear
membrane (Martin and Russell, 2003). Functional dependence supposedly
started with a strictly anaerobic archaebacterium hosting a facultative anaerobic
-proteobacterium. Endosymbiosis was established on the basis of H2 and CO2
production by the -proteobacterium. The Archaean host then metabolised the
H2 and CO2 to methane (Martin and Russell, 2003). Further support that the
endosymbiont represents the first-next possible closure comes from the nonnucleate phase the eukaryote cell passes during cell-division. This periodic loss
of the nucleus is a strong indication that the nucleus never was an
General introduction
endosymbiont,
membrane.
because
an
endosymbiont
cannot
easily
‘abandon’
47
its
Depending on the criteria applied, the endosymbiont and the nucleus can both
be regarded as the first-next possible closure. Starting with a bacterial/archaean
cell, the property of engulfment of an endosymbiont may be regarded as the
high complexity state that allows transition to the endosymbiontic state. In this
case, the hypercycle of the endosymbiont causes the structural closure in
relation to a hypercycle, and the recurrent interactions with the hypercycle of the
host cell represent the recurrent functional relationship. However, after having
transferred control over its metabolism to the host cell, one may also consider
the endosymbiont as an organelle. The endosymbiontic organelle now allows for
a high complexity cell, while the nucleus and its mediating role between two
functional levels in the cell represent structural and functional closure. As the
repeating of internal closure levels is not accepted by the operator theory, the
choice for either one of these situations makes the other obsolete and will not
affect the overall structure of the operator hierarchy.
In practice the coupling between the endosymbiontic relationship and nucleus
formation is so strong that cells with endosymbionts that lack a nucleus play no
role in existing phylogenies. On this basis, this thesis refers to endosymbiontic
cells as ‘endosymbiontic/eukaryotic’ or as ‘eukaryotic’.
Figure 1.15. Formation of an endosymbiontic/eukaryotic cell resulting from the
combination of a mediating membrane and a
second-order catalytic set of molecules
The bacterial or archaean multicellular
The literature offers various examples of non-endosymbiontic ‘multicellularity’
(Fig. 1.16) including Myxobacteria, Actinomycetes and Cyanobacteria (the bluegreen algae) (e.g. Bonner, 1998). When considering multicellularity in the
context of the operator theory, we should search for a new mechanism
supported by the cell, because this mechanism will allow for a new closure type.
The most complex state of a bacterial/archaean cell shows internal
compartments surrounded by membranes and can interact with other cells by
means of plasma strands. Based on these complexity criteria, the operator
theory regards functional differentiation of colonial cells not a sufficient criterion
48
Chapter 1
for multicellularity. In other aspects, the discussion of prokaryote multicellularity
does not differ much from the discussion of eukaryote multicellularity. For this
reason prokaryote multicellularity is not addressed in more detail.
Figure 1.16. Formation of a multicellular operator
resulting from recurrent bonds between cells and
recurrent catalytic dependence
The eukaryote multicellular
Starting from eukaryote cells, the first-next possibility for closure is that in which
interacting cells create a multicellular organism (Fig. 1.17). Independent
development of multicellularity in many branches of the phylogenetic tree
indicates the existence of powerful evolutionary forces in the direction of cell
cooperation. A basic reason for this lies in intrinsic advantages, such as higher
enzyme concentrations for digesting food, larger size and the three-dimensional
use of space. Once multicellularity has developed, forces between interacting
multicellular units will represent the dominating selective factors.
What exactly defines a multicellular organism? To be recognised by the operator
theory as a unit system, the interactions between the cells in any multicellular
should create a new emergent property that unifies the interacting cells both
functionally and structurally. To become a structural whole requires structural
closure, which means that recurrent structural bonds have to be created
between the cells. Functional closure, then, requires recurrent interactions. This
is fulfilled by mutual autocatalytic dependence of the cells. Based on this
reasoning, I propose that multicellular operators can be defined as a group of
cells that show a combination of structural linkage and functional interaction
causing at least one recurrent process that, under ambient conditions, is
obligatory for the autocatalytic functioning of the contributing cells. In cases
where cells interact structurally and functionally but do not have to interact
recurrently for their autocatalytic functioning, the cellular unit must be considered
General introduction
49
a multicellular colony. Many multicellular systems, including mammals, pass
through a single-celled stage (the zygote) and multicellular colonial stage (for
example, the two-, four- and eight-celled stage of a human embryo) before they
develop into a multicellular organism.
Figure 1.17. Formation of an endosymbiontic/
eukaryotic multicellular operator resulting from
recurrent bonds between eukaryotic cells and their
recurrent catalytic dependence
Several reasons explain why the above general definition of multicellular
operators allows many different system types. Firstly, cells can be connected in
several ways. Secondly, cells may functionally interact in several ways. Thirdly,
interactions between cells may be based on different combinations of structural
and functional possibilities. Finally, as soon as organisms consist of many cells,
the interactions between the cells may create hierarchies of internal processes
that may range from local to general.
An aspect that is still not covered by the above definition is that interactions
between cells may lead to different ‘cooperation intensities’ that lay between the
following extremes: ‘brittle’ interactions and intense communication.
Relatively 'brittle' interactions are formed when the cells are bound in the least
forceful way by membrane proteins and when the diffusion of proteins conveys
the intercellular communication. In this case, the overall system is hardly more
than a colony. The slug-like structure plasmodium illustrates this. A plasmodium
is formed when many individually dwelling and genetically different cells of slime
mould come together. The cells in the plasmodium are kept together by
membrane proteins and communicate with each other through chemical signals.
Together the cells produce higher concentrations of digestive enzyme, ensuring
a better survival under low food conditions. They also reproduce together. To
reproduce, they must depend on the multicellular stage to initiate differentiation
and the construction of a stalk on which spores are formed.
50
Chapter 1
At the other end of the spectrum, we find cells that intensively communicate
through plasma strands. Such plasma connections are allowed through gap
junctions between animal cells, through plasmodesmata between plant cells,
and through incomplete cell walls such as in some fungi or developing insect
eggs. Multicellularity requires, however, that only the plasma be connected,
while the genetic material is retained in the cell.
From the point of view of first-next possible closure, the operator theory expects
a new type of structural and functional closure as the basis for multicellularity.
Because individual cells already can show chemical communication,
multicellularity should involve a new type of intracellular communication. The
most likely candidate for this are the above mentioned plasma connections
between cells. Accordingly, the use of closure dimensions in the operator theory
suggests that cells in a multicellular organism typically must communicate by
means of plasma connections between their cells.
Several aspects in relation to the above definition of multicellularity require some
discussion.
Firstly, although the above definition does not exclude an organism consisting of
cells with different autocatalytic sets, genetic conflicts about which genetic
material participates in reproduction will generally select against chimeras.
Secondly, the success of a multicellular organism in maintaining its organisation
depends on its capacity to deal with prevailing environmental conditions and to
adapt or to repair its organisation after it has been damaged, for example, the
re-growth of a twig from a willow tree that has been stuck in the ground. If the
weather is not too warm, too cold, or too dry, etc. and the stick has enough
buffered resources, it may survive long enough to regenerate roots and start
growing, and eventually become a willow tree. In extreme cases, even single
cells of an organism can be induced to regenerate the multicellular structure.
With a little help, for instance, by in vitro-culture, this applies to many plant cells.
The callus that forms and the leaves and small plants that may emerge from it
can be considered multicellular operators in the in vitro environment. Yet, they
need roots and a certain size and strength before they have assembled all the
tools needed for internal interactions to survive as a plant in an outdoor
environment where they can develop towards a reproductive stage. The
example of a twig shows that the interaction between the environmental
conditions and the state of any particular multicellular organism determines
whether a colony of cells can regain organism structure. The above definition of
multicellular operators is explicit in recognising this relativity.
General introduction
51
Thirdly, lichens and plants with mycorrhiza are 'multicellular organisms'
consisting of two or more closely interwoven types of cells that belong to
different species. As lichens frequently disperse in the interwoven form, it seems
as if their reproduction is based on a mixture of cell types. However, lichens can
be classified as a symbiosis because the fungi show sexual reproduction without
the algae’s genetic involvement and can survive, at least temporarily, as
individual multicellular organisms during this stage.
Finally, a multicellular organism may produce structural material that supports its
growth. This material is generally more persistent than the organism. For
example, a flash of lightening may split a tree’s bark and root system into two
separate halves. These two halves may continue to grow. Although the two tree
halves are still connected through the wood, they show no intercellular
communication anymore and have become separate organisms.
The first-order neuron cycle
Because adding more cells to a multicellular organism does not affect its closure
type, multicellularity forms a serious complexity barrier. The operator theory now
suggests a search for the most complex property that interacting cells allow and
that represents the shortest route to a novel functionality. If chemical
communications between cells by diffusion gradients (as already existed as the
basis for multicellularity) is compared with communication by cell-specific
contacts over large distances that neural extensions of cells allow, the latter
represents a new interaction and introduces a higher complexity interaction type.
These qualities make it the preferred choice in the context of the operator
hierarchy.
Figure 1.18. The figure shows a first-order
interaction cycle based on neuron interactions.
These types of modules of interacting neurons
function as the most basal neural memory and are
Called Categorising And Learning Modules
(CALMs).
52
Chapter 1
Based on interactions between nerve cells, groups of cells with closed
excitation-inhibition dynamics can be created (first-order interaction cycle) (Fig.
1.18). Such groups of interacting nerve cells or 'modules' can exhibit special,
novel dynamic properties. One example of such new dynamics is the
categorising and learning function performed by the neocortical minicolumns.
Another example is the oscillation pattern of 'interior olive' modules in the
cerebellum, which function as neural metronomes (McCormick, 1995). Several
authors have advocated that such first-order interactive modules are the brain’s
basic functional units (Mountcastle, 1975; Szentagothai, 1975). In the following
text I will refer to first-order neural modules as 'Categorising And Learning
Modules' or CALMs, a name that Murré (1992) and Murré et al. (1989, 1992)
introduced for computer simulations of the neocortical mini-columns.
The neuron hypercycle
A disadvantage of the single CALM is that an earlier learned task is easily
forgotten during the training of a next task. To overcome this problem, Happel
(1994) has examined the possibility of coupling CALMs in a recurrent way (Fig.
1.19). Due to the circular stimulation of CALMs in a recurrent network, the
interaction strengths of the connections between the cells in the CALM network
adapt until they reach a stable state that supports dynamic patterns of nerve
firing. These dynamic patterns can be periodic or chaotic. The transitions
between periodic and chaotic behaviour are relatively abrupt. Such abruptness
agrees with the almost discrete limits found in the state space of the neural
network dynamics. These limits separate observations in what appears as
distinct classes. The inclusion of a temporal aspect implies a large extension of
state spaces compared to any first-order recurrent, non-hypercyclic network. As
Happel (1997, PhD thesis chapter 4: ‘Modular functions’) shows, an interesting
property of recurrent networks is that they can capture the complexity of newly
learned tasks with little loss of earlier experiences.
The emergence of the neural hypercycle is the second time that an operator
emerges as a consequence of an internal differentiation. The first case is, of
course, the endosymbionts in the eukaryote. The reason why this 'internal'
closure must be preferred above potential cooperation between organisms is
that the operator hierarchy is based on first-next possible closure steps.
Compared to cooperation between multicellular organisms (e.g. fungi, plants,
algae), the interactions between multicellular neural units offer a more direct
transition than those between individual organisms. The fact that the cooperation
between multicellular organisms represents a less direct route makes organism
interactions inconsistent with the operator hierarchy.
General introduction
53
Figure 1.19. A second-order neural interaction
cycle based on recurrent interactions between
CALMs
The neural hypercycle probably emerged after primitive neural webs were
formed. Several reasons explain why this sequence occurred. Firstly, the neural
hypercycle requires a minimum amount of nerve cells to form neuronal modules
and their interactions. Secondly, an organism needs 'knowledge' about its
environment to survive. Plotkin (1994) defines 'knowledge' as any kind of system
structure that enables an entity to respond to and act in its environment. Using
this definition, there is no inherent knowledge in the CALM-modules of a newly
born organism, because it has not yet learned anything. It is therefore necessary
that an animal can start its life with a more or less prewired neural circuitry that
allows for reflexes. Due to selected-for, gene-based ontogeny, these reflexes will
provide the right responses to survive and to reproduce. In a later phase the
supporting environment of a pre-wired brain will allow the emergence of brain
structure that is capable of learning. As predetermined structure also strongly
reduces the time to learn something, some predetermined structure in the brain
is also useful for organisms that have to learn. The above discussion shows that
learned behaviour, which is based on CALM-webs, must be considered to have
evolved in a context of a hardwired brain that produces most of the organism’s
behaviour that it did not have to learn.
The sensory interface
The sensory interface that surrounds all hypercyclic neural networks has never
existed as a separate entity. It co-evolved with the neural network in multicellular
organisms and provided the neural network with information about its
environment.
54
Chapter 1
The sensory interface can be divided into two parts: an observing part, the
'perception interface', and an activating part, the 'activation interface'. The
perception and activation interface are both integrated into the 'vehicle' in which
the network resides. In organisms, the perception interface includes sensors
focused on the outside world, such as eyes, ears, nose, etc., and all sorts of
proprio-receptors within the organism such as chemo-receptors, stretch
receptors, temperature receptors, etc. The activation interface includes nerves
that, for example, stimulate muscles for activity directed at the outside world and
stimulate the neuro-secretory cells for internal regulation of physiological
processes.
The memon
The memon is created by a combination of two closures: a closure creating
second-order recurrent interactions between neurons in a recurrent CALM
network, and a closure creating a physical limit in the form of the sensory
interface (Fig. 1.20). These two closures together allow the memon to emerge.
.
The name memon comes from Dawkins (1976) who used the word meme for
cultural replicators that thrive in brains and for the artificially manufactured
products of brains, such as books, computers, and so on. In the operator theory,
however, the word meme is used in a differentiated way: it describes more
precisely a whole range of memic entities that play a role in the functioning of
memons.
Figure 1.20. Formation of a memon resulting from
the combination of second-order neuron interactions and an interface
General introduction
55
A differentiated definition of a meme is required because memons may be
involved in replicating various memic entities (Jagers op Akkerhuis, 2001):
1. Memons may exchange ideas. We suggest using the concept 'abstract
meme' (or a-meme) for such abstractions, for example, melodies, ideas,
stories, theories, and jokes that may reside in functional meme networks and
can be transmitted from memon to memon in the communication process.
This is closest to the meme concept as introduced by Dawkins (1976).
2. Memons will actively create physical 'traces' of their thoughts, such as books,
houses, computers, radio waves, etc. These traces can best be designated
as 'physical meme models' or 'physical memes' (p-memes) because these
entities represent 'real-world/physical' representations of the abstract memes.
A-Memes also replicate through these physical models because the
observation of a nice model will inspire the observer to mimic it.
3. The concept 'coding meme' (or c-meme) could be used to define any string of
codes that contains the information to construct fully or partially the neurons
and connections of a neural network. Only if this code is supplemented with
the interaction strengths can it be used to copy fully the network’s knowledge.
In fact, c-memes code for a neural network architecture and associated
knowledge much in the same way as genes code for the chemistry playing a
role in the survival of cellular beings.
As far as I know, the above way of differentiating between various memic
entities does not yet play a role in modern 'memology' (e.g., Wilkins, 1998;
Blackmore 1999).
In principle, memic neural networks are not necessarily bound to organisms but
can be housed in any kind of vehicle. A memon may even be built from material
other than organic cells. In fact, any material and any other signal transmission
than electrical may be used. What counts is that the dynamic processes in a
memon allow an equivalent of the neural hypercycle with interface.
Is the operator theory a falsifiable approach?
A good theory should be falsifiable. Because the operator hierarchy rests on
mechanistic closure processes for every step in its construction and is not just a
superficial description of some accidentally ranked system states, it should be
possible to point out certain aspects of the theory that are open to falsification.
This falsification process can be split into the following three questions that
coincide with major aspects of the structure of the operator hierarchy:
56
Chapter 1
1. Can it be proven that the individual steps in the operator hierarchy are the
first-next possibilities for the suggested closures?
The operator hierarchy can be tested step-by-step to see whether it is based
on existing system types and realistic pathways for their creation. This control
method focuses on the physical existence of the system types in the operator
hierarchy and on proof for the transitions through which they have been
formed.
For most of the system types in the operator hierarchy, the physical existence
is part of generally accepted scientific knowledge. Apart from philosophical
discussions about how to prove that anything exists, little scientific doubt
exists about the measurability of entities such as atoms, molecules, cells, etc.
The fact that 'there really is something out there' that corresponds with our
mental models is supported by the increasing accuracy with which these
systems and their individual and interactive properties can be described and
by the fact that the use of different ways of observation yield identical results
(Kragh, 1999).
In this context the mechanisms have already been discovered by which
relatively simple systems can be artificially recreated from lower level building
blocks, for example, by creating hadrons from nuclear particles, atom nuclei
from hadrons, atoms from nuclei and electrons, and molecules from atoms.
Higher level systems have not yet been recreated by de novo selforganisation. Yet, I am confident that with increasing scientific insight and
technology, the day will soon come when the cell will also be added to the list
of systems for which it will be possible to recreate the appropriate selforganisation conditions.
All minor transitions in the operator hierarchy can be tested to see whether
they comply with the concept of first-next possible closure. Testing this
requirement is simple. One need only to show that at any position in the
hierarchy, the present logic is incorrect and needs to be improved by adding
a step that currently is not part of the hierarchy or by removing a step that
currently is included in the hierarchy. If only a single system can be added or
removed, the structure of the operator hierarchy in its present form must be
considered invalid. This would lead to the need either to improve the operator
hierarchy or, when this is not feasible, to reject it.
2. Can the re-arrangement of the operators be tested to see whether they follow
a coherent logic according to closure dimensions?
The rules for recognising closure dimensions can be tested. For example, the
interface closure dimension should successfully apply to all operators that it
characterises. These operators are the fundamental particle, the quark
confinement, the electron shell, the cell membrane and the sensory interface.
The validity of closure dimensions can furthermore be tested by their
extrapolation and subsequent use for predictions of future system types. This
point refers to predictions of the emergent properties of several not yet
General introduction
57
existing system types with a higher complexity than the hardwired memon
and to the prediction that the present ‘fundamental particles’ (e.g., quarks)
are the lowest level of closed, finite structure. Of course, the prediction of a
system type that does not yet exist is difficult to falsify. The time may soon
come, however, when complex memic systems will be constructed. At that
future moment, it will become possible to show that, for example, hypercyclic
technical neural circuitry must indeed form the basis for technical intelligence.
Similarly, it can be shown that these future memons use structural autocopying (the copying of the neural network) as a simple means to provide
their offspring with the knowledge they need to live and to survive. The
prediction that fundamental particles form the least complex, basic system
type on which the operator hierarchy is founded would become testable with
increasing insights into the lowest levels of matter’s organisation. If a still
lower level of organisation of matter based on structurally closed units could
be proven to exist, this would lead to adaptation or rejection of the current
structure of the operator theory.
3. Is it possible to falsify the operator hypothesis? In other words, can the
suggestion be proven or disproven that nature’s possibilities for constructing
closed system topologies are so limited that the currently identified operators
and closures indeed represent the only options?
Answering the latter question requires insight into the whys of all details of
the operator hierarchy’s structure. However, the most important question that
needs to be answered is why every primary operator introduces a new
closure dimension. Such insight is not only needed to understand the
operator hierarchy, but it also offers pathways to falsify the operator
hypothesis; a fundamental understanding may indicate that there are other
ways to create closures/operators but these are not included in the present
theory. Currently, this subject is unknown territory and offers a major
challenge for future research.
In conclusion
Various fundamental aspects of the operator theory can be refuted in some way
or another. Consequently, the theory represents a refutable construction.
Thermodynamics and stability
It is the author’s conviction that a general approach to system hierarchy should
fully agree with thermodynamic laws but should not be distracted by
thermodynamic justification from its main theme: the hierarchy of organisation.
No system in our universe has ever been found to violate thermodynamics on a
large scale and for a long time, or when it does, any increase in internal
58
Chapter 1
organisation is compensated by a bigger increase in disorder outside the
system’s limits (Prigogine & Stengers, 1984). Since no deviations have ever
been demonstrated from this rule, it is a rather theoretical exercise to demand
proof that any existing system does not violate the laws of thermodynamics. A
consequence of this reasoning is that self-organised system states must
represent a thermodynamically advantageous phenomenon. Otherwise, nature
would never have been able to repeatedly produce complex systems from less
complex elements. In short, as long as we limit a general system hierarchy to
systems that exist in our universe, we may take the thermodynamic justification
for granted and focus on structural aspects of complexity instead. When
predicting future system states, however, these states must be checked for
consistency with thermodynamic laws.
In addition to thermodynamics, stability is another factor determining the
presence of systems. The reason is that systems will be selected away when
they are unstable as the result of a problematic internal organisation or as the
result of disintegrating interactions with the environment. Only those systems
that combine an easy production with a long enough ‘shelf life’ prevail in the
evolution of system types. This is, for example, the reason why hydrogen is so
abundant in the universe or why the light and easily formed pions convey the
force between hadrons in the atom nucleus. And if an operator shows
reproduction, the production of offspring allows it to escape the limits of its
proper existence and structural stability by creating one or more copies before it
dies.
Neither entropy production nor production of higher level particles is
homogenously distributed in the universe. Once organised structures have
formed, increasingly complex structures start to emerge at an increasing rate.
Kurzweil (1999) has integrated these observations into his ‘Law of time and
chaos’, which states that ‘In a process, the time interval between salient events
(that is, events that change the nature of the process, or significantly affect the
future of the process) expands or contracts along with the amount of chaos’.
Thus, in a chaotic environment, salient events occur at a low rate, and when
order increases, this rate increases. Assuming that first-next possible closures
can be regarded as salient events, Kurzweil’s viewpoint fully merges with the
operator hierarchy.
General introduction
59
Outline of the thesis
The major aim of this thesis was to search for rules that guide the development
of structural complexity in the universe and that can be used to create unity in
science by improving the analysis of hierarchical organisation in biology, ecology
and other sciences.
In relation to these goals, basic principles of physical and ecological hierarchies
were studied. The need to follow a fundamental approach was based on the
observation that these hierarchies were generally flawed by imprecise and/or
overly general rules for selecting and ranking elements.
The subsequent search for a methodology that would allow a structured analysis
of hierarchy was based on two premises: (1) all the elements in a strict hierarchy
must have their proper positions, such that elements can neither be removed nor
added without corrupting the hierarchy, and (2) it is possible to find a ranking
rule to create a strict hierarchy for all particles in nature, regardless of whether
these particles are abiotic or biotic.
It is thought that the value of the thesis lies in the fact that identifying a strict
hierarchy is of general importance for science for two reasons. Firstly, it can
have a restructuring influence whenever existing practice is flawed by less than
strict hierarchical rankings. Secondly, a strict framework for the hierarchical
ranking of operators can be used as a strong instrument to integrate different
fields of science. The product of this thesis, therefore, must be considered of a
qualitative rather than quantitative nature, in the sense that the final aim is to
contribute to the structuring of scientific thought. In relation to the above goals,
the subjects discussed in the chapters in this thesis are arranged more or less in
order of increasing generality and from fundamental to applied topics.
Chapter 2. The birth of an idea
This chapter presents the basic principles underlying the operator hierarchy. The
text elucidates the choices that were made to overcome major difficulties when
deciding which elements should be included in the hierarchy and which should
not, and the formulating of strict rules for ranking. In this chapter general rules
for evolution are deduced. These rules can be applied to all operators,
regardless whether they are particles or organisms.
Chapter 3. The future of evolution
Predicting is difficult, especially predicting the future. Yet, analogous to using the
periodic table of elements to predict new atoms, the strict structure of the
operator hierarchy can be used to predict new operators. All the operators
predicted in the nearby future will be technical neural network organisms. The
realisation that neural network organisms cannot use genes to copy their
60
Chapter 1
knowledge forms the basis of a critical appraisal on the lack of parallelism
between the concepts of genes, representing autocatalytic information, and
memes, defined by Dawkins as cultural replicators.
Chapter 4. A backbone for system science
This chapter demonstrates how the operator hierarchy can be used to organise
various aspects of natural organisation. It is advocated that the hierarchical
relationships between the operators be ranked in a strict way, but that the
situation is different within operators and in systems consisting of interaction
operators. With respect to the latter situations, a new methodology is introduced
that looks at hierarchical relationships from four different points of view:
displacement, information, construction and energy (DICE).
Chapter 5. On life and death
The definition of life has long been a major puzzle in science. The difficulty of
reaching a consensus may have various causes, including: (1) the difficulty of
deciding on the minimum requirements that organisms must show to be
considered living beings; (2) the fact that many definitions include facultative
aspects, such as reproduction or metabolism; (3) the confusion about the
concepts of life, as a property of matter, and living, as the dynamic activity of
systems representing life, while there is also confusion about the use of life as
the ensemble of all life on earth in contrast to life as a property relating only to
organisms; and (4) the confusion about whether a mechanistic explanation of
the cell offers a sufficient basis to define both bacteria, protozoa, plants and
animals. In relation to the latter points, the operator hierarchy has the advantage
of offering a completely new, hierarchical context to deal with problems of
definition, which -- mutatis mutandis -- also helps to solve definition problems
with respect to the concepts of the organism and of death.
Chapter 6. Getting the biggest part of the pie
Although Darwin talked about selection acting against the less well endowed,
the evolution process became known as the struggle for life and survival of the
fittest. Fitness has become related to the number of successful offspring. For
various reasons, this represents a limited way of measuring evolutionary
success. First of all, this idea makes it impossible to measure the success of
units of evolution during periods when they are not reproducing, for example,
when they are involved in non-reproductive processes, such as competition
through subterranean shoots or the growth of a colony. Furthermore, abundance
is not always an accurate predictor of the impact that offspring can have on
ecosystem processes. To contribute to this topic, it is suggested to focus on the
control that evolutionary units have over resources in their environment. In the
same way as dimensions for hierarchy have been used in the operator theory,
dimensions for resource dominance have been proposed for analysing
evolutionary success.
General introduction
61
Chapter 7. Towards consistent definitions for the operator theory
Although the operator theory represents a straightforward reasoning that
focuses on how self-organisation steps produce a long sequence of operators,
the reasoning behind this simplicity is quite intricate. To define as clearly as
possible the logic of the operator hierarchy, this chapter discusses the definitions
of the concepts used in the operator theory.
Chapter 8: Centripetal science: contributions of the operator hierarchy to
scientific integration
The main goal of this thesis was to develop an integrating theory that can
counteract the forces of hyperspecialisation and compartmentalisation resulting
from the knowledge explosion. The most important result of this goal was the
development of the operator theory. This chapter discusses the contributions of
the operator theory to several well-known integrating theories, such as a cosmic
timeline, ecological hierarchy, periodic tables, Darwinian evolution and also
discusses a meta-level analysis of how unifying concepts can be linked to the
levels of the operator hierarchy.
Chapter 9: General discussion and conclusions
This chapter places the conceptual framework of the operator hierarchy and its
applications in a broader context. This process is guided by the search for
answers to the following five questions: (1) what kinds of novel theoretic
developments have accompanied the elaboration of the operator hierarchy, (2)
what impacts may the operator theory have on science, (3) what aspects of the
operator theory require discussion because of potential weaknesses, (4) how
can the operator theory be evaluated, and (5) what routes are open for future
development.
62
Chapter 1
Chapter
2
The birth of an idea
Operators, the Lego-bricks of nature. Evolutionary
transitions from fermions to neural networks
G.A.J.M. Jagers op Akkerhuis and N.M. van Straalen
World Futures, The journal of general evolution 53: 329-345, 1999
What seems to physicists to be a hopelessly complicated process
may have been what nature found simplest,
because nature could only build on what was already there.
(Francis Crick)
64
Chapter 2
Abstract
When Darwin wrote his ‘On the origin of species...' (1859) he focused on
evolution as a property of living organisms in interaction with abiotic and biotic
elements in the world. This viewpoint is still dominant amongst biologists. For
particle physicists and cosmologists evolution refers to a larger scale, ranging
from quarks and atoms to galaxies, stars and planets (i.e. Pagels, 1985,
Hawking, 1988). To close the gap between such different viewpoints, a wide
range of perspectives on an interdisciplinary understanding of system
development has been published (i.e. Teilhard de Chardin, 1966, 1969, von
Bertalanffy, 1968, Varela, 1979, Prigogine and Stengers, 1984, Laszlo, 1996).
As an integrative concept, the construction of nature from a hierarchy of system
layers forms a central tenet in general system research and the stepwise
construction of this layered hierarchy can be regarded as an interdisciplinary
evolution theory. Surprisingly, the literature offers no unequivocal rules to
recognise a multilayer hierarchy in nature. This presents an obstacle for
interdisciplinary approaches to evolution.
Searching a solution to part of the above hierarchy problem, the present paper is
dedicated to the analysis of a special kind of layering in natural systems, which
is based on transitions between ‘building block' systems. To identify these
building blocks, and the transitions from building block x at level A, to building
block y at level B, the focus of this study is further limited to 'hypercyclic
dynamics' and ‘containment'. On the basis of these criteria, a hierarchy is
created which shows no possibilities for ‘bypasses'. It connects hadrons to
atoms, atoms to cells, and cells to neural networks. Implications of this hierarchy
for system studies and evolution are discussed.
The birth of an idea
65
Introduction
System thinking has opened up many ways for the examination of systems, their
internal organisation and their external relationships. This has led to general
laws on system organisation on the one hand, and a clearer view on the
differences that exist between system types on the other.
A general aspect of all system studies is that reality is regarded as to show a
layered structure, which is minimally represented by a system and its elements.
Stressing the importance of hierarchy in science von Bertalanffy (1968) wrote
that ‘A general theory of hierarchic order obviously will be a mainstay of general
systems theory'. That hierarchy is omnipresent in science is reflected in the
many metaphors which have been proposed for it, including the ‘worlds within
worlds' approach, which according to Close dates back to the Japanese
physicist Kaku (Close, 1983), the ‘cosmic onion' (Close, 1983) and the ‘Chinese
boxes' (i.e. Simon, 1962, Koestler, 1967, Laszlo, 1972). A hierarchy based on
unit systems, which are characterised as being ‘formed' and ‘centred', has been
proposed by Teilhard de Chardin (1966). Still other studies have explored the
mathematical formalism of layered structures as consisting of units composed of
interacting elements (Goguen and Varela, 1978, Geiger, 1990) or, with respect
to ecosystem interactions, in the form of a 'biomatrix' (Jaros and Cloete, 1987).
The presence of hierarchy in different areas of system research can furthermore
be inferred from the use of concepts such as transformation, emergent
properties, the top-down viewpoint of reductionism, and the bottom-up
viewpoints of holism and constructivism, the occurrence of transitions, symmetry
breaking, bifurcation, attractor states, integrated elements building 'holons',
autopoiesis, etc. (i.e. Feibleman, 1954, Koestler, 1978, Varela, 1979, Labson,
1985, de Kruijff, 1991, Beloussov, 1993, Laszlo, 1994, Capra, 1996, Szathmáry
and Smith, 2002).
Interactions which cause new system layers have been regarded as ‘quanta of
evolution' (Turchin, 1995). The quantum aspect of these transitions results from
an all-or-nothing restriction of processes in the original system, which creates
new structures and associated dynamic properties (Heylighen, 1995, 2000). For
example, the cyclic restriction of enzymatic interactions and a confining cellular
environment are required before the arisal, or ‘emergence', of reproduction is
possible and a new layer in the hierarchy can be recognised.
The analysis of hierarchy forms the main topic of the present study. As a point of
special interest, we investigate the possibility for a general, yet strict hierarchical
classification of special ‘building block' systems. For any hierarchy, we consider
66
Chapter 2
the construction of sound layers a necessity to avoid that the layering is
corrupted via ‘bypasses'. An example of what we consider a corrupt hierarchy is
the sequence planet-stones-sand. It is perfectly possible to construct a planet
from sand alone, and in this way bypass the ‘intermediate layer' of the stones. In
robust hierarchies such bypasses do not exist and complexity can be ranked
solely in a strict layer-by-layer fashion.
In the present search for a rigorous hierarchy in nature we deliberately restrict
our efforts to ‘building block systems' or ‘unit systems'. The idea is, that the use
of a kind of natural 'Lego-bricks' allows a dissection of system complexity in
stepwise transitions from building block x at level A to building block y at level B,
etc. As we will explain below, the building blocks that can be identified in this
way include the hadrons, the atoms, the molecules, the cells, the multicellular
organisms and a special kind of neural networks. All other natural systems do
not fulfil the present building block definition. Instead, they are regarded as
‘interaction systems' and enter the present discussion only on special occasions,
for example when we discuss hypercyclic interactions and their containment.
Examples of interaction systems include stars, planets, ecosystems, society, a
football, etc. The present focus on building block systems requires, however,
that we define clear criteria to recognise them.
As the criteria for building blocks we used hypercycle formation as the primary
aspect, and different forms of compartmentation as the secondary aspect.
Hypercycle formation. Elements which perform a cyclic process can interact to
create a secondary reaction cycle. Such a ‘cycle of cycles' is called a
‘hypercycle' (Eigen and Schuster, 1977). Hypercycles have highly special unitproperties. The enzymatic hypercycle, for example, makes reproduction possible
(Eigen and Schuster, 1977, Eigen, 1985, Kauffman, 1993). A schematic
representation of an enzymatic hypercycle (Eigen and Schuster, 1977) is shown
in Fig. 2.1.A to illustrate the close relationship between structural and functional
aspects of such systems. A less abstract version (Kauffman, 1993) is shown in
Fig. 2.1.B to indicate that actual physical hypercycles look rather like ‘webs' of
interactions without that the central hypercycle can be recognised structurally.
Compartmentation. Within each large group of systems based on the same type
of hypercycle, the mechanism of compartmentation is used to recognise internal
subsets. The most fundamental kind of compartment formation involves the
containment of a hypercyclic set of elements by a layer, or ‘interface', which
mediates the interactions between the elements of the hypercycle and the world.
As such, it offers a natural system limit for thermodynamic considerations. An
example of an interface are the electron-wave clouds which surround nuclei and
mediate interactions with neighbouring nuclei. A different way of compartment
formation is observed when two or more systems with a contained hypercycle
interact to form a multiplet structure, for example when atoms interact to form
The birth of an idea
67
molecules. As we will discuss below, the mechanisms of hypercycle formation
and subsequent compartmentation can be used to create an unambiguous
hierarchy of building block systems.
For reasons of clarity, the term ‘operator' is introduced as a common name for all
the building block systems which consist of a contained hypercycle, and the
systems which are multiplets hereof. Accordingly, the present approach is
regarded as the ‘operator approach' or ‘operator framework'. The recognition of
operators as a special group of systems has several advantages. First, it helps
to distinguish between operators, the building blocks, and other systems, which,
as was discussed above, consist of interacting operators without being operators
themselves and were called ‘interaction systems'. Another advantage of using
the operator concept is that it separates operator evolution from biological
evolution, biology being limited to the subset of operators based on cells and the
forces which cause diversification and selection at that level.
Although this study focuses on structural aspects of systems, these are
considered as the inseparable mirror image of the underlying dynamics. The
mechanism for all dynamics lays in entropy increase. For discussions of the
application of the laws of entropy to non-linear systems is referred to studies by
Schrodinger (1944), Prigogine and Stengers (1984), Eigen and Winkler (1983),
Schneider (1988), Schneider and Kay (1994), Swenson (1989) and others. For
cellular and higher levels, these studies have shown that these systems can
organise themselves and reduce their internal disorder, which creates an
entropy decrease, as long as this is ‘paid for' by an equal amount of entropy
export to the outside world.
In the below text we start with an inventory of all operators in nature, on the
basis of an analysis of their construction via hypercycle formation and different
forms of containment. These operators will be ranked in a hierarchical
sequence, which is subsequently analysed for possible internal structures. At the
end of this study, we will discuss the present system approach in relation to an
interdisciplinary viewpoint on evolution.
68
Chapter 2
E1
A
E2
En
E4
E3
Figure 2.1. Two representations of an enzymatic hypercycle. Part A shows a more
abstract, cycle oriented representation, which can be found in the work of Eigen (1977).
Here E1 to En represent enzymes. Part B shows a network representation of cyclic
enzymatic processes (after Kauffman (1993) ‘The origins of order'. With permission of
Oxford University Press). Essential of both graphs is that the enzymatic reactions in
themselves form cyclic events, which via their linking in an overall cycle have become
functionally unified into a catalytic hypercycle.
The birth of an idea
69
Three similar construction sequences
To elucidate the way in which we analyse steps between operators of different
complexity in order to create an inventory of all operators in nature, the path
from the atom to the cell is shown as an introductory example (Fig. 2.2). In the
first step atoms link to form atomary multiplets, the molecules. From a
topological viewpoint the structure of molecules is special. The reason is that no
matter how many atoms are added and no matter how complex threedimensional structures are built this way, any molecule remains a system of
connected atoms showing the multiplet structure. An escape from this limitation
requires another interaction than atomary linkage, and thus a new emergent
property. This is allowed by catalytic interactions in which molecules, for
example enzymes, transform substrate molecules. In special cases, the product
of a catalytic process equals the catalyst of a next, etc., which may cause a
cycle of catalytic reaction cycles, called a ‘hypercycle' by Eigen and Schuster
(1977, 1978a, 1978b). The mechanistic basis hereof is explained in great detail
by Eigen and Schuster (1979), Eigen (1985) and Kauffman (1993). The catalytic
hypercycle performs a new dynamic property, that of 'autocatalysis', normally
referred to as reproduction. As long as the set of autocatalytic enzymes lacks a
boundary, or ‘interface', it can not be considered an operator. If the interface is
formed by a molecular membrane, we can regard the resulting unit a living cell.
The formation of the cell thus requires the simultaneous occurrence of two
emergent properties; hypercyclicity and containment by a membrane. The
requirement of the simultaneous occurrence of these properties favours
explanations of contained autocatalysis from a ‘seed' instead of from a ‘soup',
which possibilities were recently discussed in a review of theories about the
origin of the cell by Edwards (1998).
70
Chapter 2
Figure 2.2. Example of a basal pattern in operator evolution: interactions leading from
atoms to cells. The steps being explained in more detail in the text, the following phases
can be distinguished. Atoms combine to molecules, some of which can act as enzymes.
Sets of enzymes may show catalytic closure and form an autocatalytic hypercycle. When
this is surrounded by a membrane, the cell is ‘born'. Phases during which system types
are not considered operators are shaded.
As is shown in Fig. 2.3, two additional construction pathways, similar to that from
atom to cell, can be recognised in the range of all operators known to science.
The first describes the sequence from quarks to atoms, the other that from cells
to neural networks.
The mechanisms behind the first sequence are being studied by particle
physicists. The most fundamental particles that science has experimental
access to are the particle-like Dirac-fermions, which include the quarks and
leptons, and the force carrying bosons, such as photons and gluons (Close,
1983, Feynmann, 1985, Pagels, 1985, Kaku, 1987, Witten, 1988, Hawking,
1988, `t Hooft, 1992, 1994, Wilczek, 1998). Quarks continuously emit and
reabsorb clouds of gluons which can ‘bind' the quarks forcefully into a multiplet
structure. Pairs of quarks are called mesons and triplets are called baryons. Well
known baryons are the proton and neutron. All baryons posses the special
property that they can emit and re-absorb small mesons without loosing their
triplet structure. For energetic reasons such emission-absorption cycles involve
predominantly the lightest possible mesons, the pions. Recurrent pion exchange
The birth of an idea
71
between baryons causes what is called the ‘strong force', binding protons and
neutrons into lumps regarded as nuclei, representing a novel hypercyclic
structure. When the temperature of the environment drops below 3000°K,
electrons furthermore lack the energy to escape from the electric force of the
protons in the nuclei. A cloud of orbit fitting electron ‘waves' now surrounds the
nucleus as an interface. A new operator has emerged; the atom.
Figure 2.3. Inventory and ranking of all operator systems known to science. System
types are ranked according to emergent properties as is explained in the text. Phases
during which systems types are not considered operators are shaded. The inventory
shows two marked regularities. Firstly, the sequence showing the formation of cells from
atoms can be seen to recur in the way in which atoms are formed from quarks, and
hypercyclic neural networks are formed from eukaryotic cells. Secondly, the inventory
shows an overall pattern of increasing possibilities for compartmentation and
differentiation in operators, which is further explained in Fig. 2.4.
The other sequence, at the opposite side of the operator framework, leads from
cells to neural networks. These systems are studied both by biology and by the
neurosciences. The completion of the sequence from cell to neural network has
either only been possible, or has simply developed quicker on the basis of the,
more complex, eukaryotic cells. Some prokaryotic species, for example the
cyanobacteria, have also reached primitive multicellular interactions, but only the
eukaryotes developed to multicellular life forms within which neural cells evolved
which were capable of forming units of cells showing recurrent interactions.
Modern versions of artificial neural networks, as were pioneered by Hebb (1949,,
1955), show that modules of cyclically interacting nerve cells can perform
unsupervised categorisation and learning tasks. Accordingly, they have been
called ‘Categorising And Learning Modules', or CALMs (Murre, Phaf and
Wolters, 1989, 1992, Murre, 1992). Biotic equivalents of CALMs are present in
72
Chapter 2
the form of so called minicolumns in the cerebral cortex in mammals
(Mountcastle, 1975). In a number of experiments Happel (1997) has linked
CALMs in a recurrent way and investigated the properties of hypercyclic neural
networks. Although a series of subsequent CALMs can be compared with a
multilayer feed-forward network, the recurrent coupling of CALMs would require
an endless number of linearly coupled layers. As was explained by Happel
(1997), recurrent interactions make CALM networks fundamentally different from
feed forward networks, because a recurrent architecture creates fractal category
boundaries, hereby allowing for infinitely more distinctions between input
patterns than are possible when using a linear organisation (Happel, 1997, p.
69). But an isolated hypercyclic neural network does not yet fit our operator
definition. It still lacks an interface. In the form of the neural interface offered by
sense organs and activation organs this has co-evolved in the multicellular body
as efferent and afferent extensions to the hypercyclic network. It is the
simultaneous presence of the hypercyclic neural network and the interface which
marks the operator that we call the memon. With respect to the evolution of
neural networks in animals, it should be noted that the hypercyclic network must
be considered to have evolved within the context of a non-hypercyclic brain, in
which structure and functioning had a strong genetic basis. The reason is that
the genetic control of neural architecture, and therewith behaviour, originally had
a direct survival value for the organism because it prevented low fitness during a
learning phase such as is inherent to hypercyclic functioning.
An overall pattern in system transitions
In addition to the recurrent pattern of the three sequences discussed above, the
operator framework also shows an overall pattern. This is most clearly visible
after rearranging the elements of Fig. 2.3 in a staircase-like manner, as is shown
in Fig. 2.4. Now, each new hypercycle with interface is placed at the beginning
of a row, whilst the end of each row is formed by the multiplet configuration, this
being the most complex system type which is possible on the basis of structural
interactions between the operators in any row. Fig. 2.4 shows that in subsequent
rows the number of ways in which systems can differentiate before the multiplet
stage is reached, increases with one each layer. At the quark level, the
possibilities are limited to quarks and hadrons. The nuclear level ranges from
nuclei, via atoms, to molecules. The autocatalytic level includes autocatalytic
hypercycles, prokaryotic cells, eukaryotic cells and multicellular life forms. The
scientific knowledge of the last level, that of neural hypercycles, does not yet
extend beyond the memons.
The birth of an idea
73
Figure 2.4. A ‘periodic table of operators'. The inventory of operators shown in Fig. 2.3
has been split up into segments starting with a particular hypercycle (the grey bands in
Fig. 2.3) and ending with multiplets of operators which contain this hypercycle. The
vertical axis (A) indicates the occurrence of new hypercycles. The horizontal axis (B)
indicates the different possibilities for compartmentation within each group of systems
based on the same hypercycle. The lengths of the rows illustrates the number of different
operator types possible within a layer. Quarks (fermions) and gluons (bosons) directly
form hadrons. The atomary nucleus first obtains an electron shell, which may bind to form
molecules. The autocatalytic hypercycle becomes confined by a membrane which creates
the prokaryotic cells. These may either develop directly into a primitive multicellular stage,
or differentiate further to obtain an internal compartment around the basis of their
hypercycle, and then form more advanced multicellular stages. Finally, groups of neural
cells, called CALMs, are interacting cyclically and obtain an interface of sense organs and
activation organs. This we have called the memon.
74
Chapter 2
In addition to the connections between the rows, the scheme shows logic along
the diagonals. The diagonals link system types with the same emergent
property. The most obvious emergent property is that of the multi-operator
systems, examples of which are the hadrons, the molecules, and the
multicellular biota. The next diagonal is formed by the systems preceding the
multiplets. These show a more or less complex kind of informational layering. In
atoms, it is the electron shell surrounding the nucleus which forms such a layer.
In eukaryotic cells, the informational layering is represented by the nuclear
envelope, which mediates the functional separation between the production of
RNA inside, and the production of proteins outside. On the next diagonal, we
consider the prokaryotic cell to show the emergent property of structural
information copying, and on the most recently evolved diagonal, the ongoing
internal interactions which continuously change the states of the CALMs, are
used as an argument to consider the memon to show auto-evolution as
emergent property.
The scope of the present paper is limited to a study of existing system types and
their evolutionary relationships. Although the operator framework strongly
suggests some basal aspects of these systems, predictions about more complex
systems than memons, which are not yet accessible to scientific analysis, fall
outside the present scope.
Consequences of the operator approach
If the structural and associated functional organisation of operators is referred to
as their complexity, the operator framework describes the steps via which
operator evolution created complex building blocks from smaller ones. The
requirement that smaller building blocks exist and interact before larger ones
can be constructed, implies a direction in evolution, but does in most cases not
imply a directionality in the sense that the interacting operators know in which
direction they should evolve, or that they are motivated by some kind of invisible
hand with a ‘guiding' capacity.
Below the memon stage, the operators involved have never been capable of
constructing an internal representation of their surroundings to evaluate their
actions. In contrast, memons and higher level operators are not only aware of
their surroundings, but can also understand the meta-evolutionary processes
therein. These operators, therefore, may show evolution in relation to this
insight. This renders evolution a directed process in which, however, the
unpredictability of interactions remains a chance aspect. Only the existence and
the direction of this process are open to scientific inquiry. We see no way of how
to study any possible ‘goal' or ‘meaning' associated with teleological viewpoints.
The birth of an idea
75
We emphasise that the necessity for complexity to increase between operator
stages typically applies to hypercycle formation and containment steps in the
operator hierarchy. This is by no means in conflict with the decrease that any
particular operator may show in capacities when these have lost their survival
value, for example moles losing their sight.
The patterns in Fig. 2.3 and 2.4 offer a unique possibility to study the properties
that operators need to show in order to become a link in the evolutionary chain.
Such properties can be regarded as the recipe nature has used to ‘cook'
subsequent operators. We have deduced that the following operator aspects are
necessary for any operator to act as a link in the chain:
(1) Operators must show a stable internal organisation. If the operator's
internal organisation is not stable under prevailing conditions, this will have
a short term fatal effect on its functioning.
(2) Operators (in general) must maintain integrity in interactions (in general).
This represents an extension of the survival of the fittest to all operators
below and above organism-level. Of course the laws which govern
evolutionary success vary between layers, the requirements to animals in
ecosystems being of a rather different kind than to elementary particles in a
newly developing universe.
(3) Operators must be able to interact with each other and form systems which
allow for the creation of more complex operators1. If, for example, at any
place and time in the universe the most complex multi-operator does not
give rise, in the system that it is part of, to the formation of a new
hypercycle, this represents a local end to evolution. The third aspect is a
unique result of a between-operator viewpoint on evolution. It cannot be
discovered by any approach which focuses on evolution of operators within
their class.
In the light of the chaotic system that the universe seems to be, it is surprising
how rigid a backbone for evolution is suggested by the operator framework. This
rigidity is caused by the limits that emergent properties set to the formation of
new system types. There remains much freedom, however, for the actual form in
which any particular system is realised, and the moment and place in the
universe where it will occur. This freedom increases with increasing complexity
of the operators. There exist relatively few elementary particles, many atomary
nuclei, very many autocatalytic sets and an unimaginably large number of neural
network topologies. The sequence of increasing complexity operators is directly
linked with chronology (see also Teilhard de Chardin 1969, Pagels, 1985)
1
While writing later publications, the picture of analysing hierarchy using three
dimensions (interaction systems, internal differentiation and operators) emphasised that
also internal differentiations may allow for the formation of more complex operators, for
example the nucleus in eukaryote cells and the neural network in memons.
76
Chapter 2
because emergent properties of any operator are always preceded by the
operator not showing this property, or by interactions between lower level
operators in a parental system.
In conclusion
The principles discussed above allow a ranking of the building-block systems
underlying all other systems in the universe. This ranking is based on emergent
properties. The marked regularity of the resulting classification seems to indicate
that nature has little choice with respect to the kind of steps it can make between
system types. Apparently, the only freedom it has, is to let chance determine the
exact players in the game, and the moments and places of the transitions.
The mechanisms behind most of the binary steps in the operator scheme are in
principle known to the separate branches of science dealing with these systems.
The overall regularity of the scheme, however, can be regarded as an
elaboration of the cosmic onion approach by a more precise indication of the
layering of nature.
Regularities in specific groups of operators have helped to unravel underlying
mechanisms in different realms of science. Examples hereof are the 'eight-fold
way' for quarks (Gell-Man and Neeman, 1964) and the periodic table of atoms
(Mendeleev 1871). In analogy, we expect that the periodic structure of the
overall operator scheme suggests an underlying logic. The main aspect of this
logic is the sequence of hypercycles showing increasing complexity from the
quark, via the nucleus and the autocatalytic set, to the memic hypercycle. The
formulation of the algorithm connecting these hypercycles is the closest we
think, one can come to an inclusive viewpoint on operator evolution, covering the
whole range from quarks to neural networks, and possibly beyond. Formulating
this algorithm in more detail presents a challenging field for future research.
Acknowledgements
The authors thank the following persons for stimulating discussions and/or
commenting on earlier drafts during the development of the above theory: J.
Axelsen, W. Bakker, C. Damgaard, B. Happel, P. Henning Krogh, F. Jagers op
Akkerhuis, H. Lakkenborg, H. Løkke, V. Simonsen, K. Skovbo Jensen and E.
Wattel. M. Bayley is thanked for improving the English. This paper was
supported in part by the project ‘Complex Systems' of the Danish Ministry of
Environment and Energy.
Chapter
3
The future of evolution
Extrapolating a hierarchy of building block
systems towards future neural network
organisms
G.A.J.M. Jagers op Akkerhuis
Acta Biotheoretica 49: 171-189, 2001
The future belongs to those who believe in the beauty of their dreams
(Eleanor Roosevelt)
78
Chapter 3
Abstract
Is it possible to predict future life forms? In this paper it is argued that the answer
to this question may well be positive. As a basis for predictions a rationale is
used that is derived from historical data, e.g. from a hierarchical classification
that ranks all building block systems, that have evolved so far. This classification
is based on specific emergent properties that allow stepwise transitions, from
low level building blocks to higher level ones. This paper shows how this
hierarchy can be used for predicting future life forms.
The extrapolations suggest several future neural network organisms. Major
aspects of the structures of these organisms are predicted. The results can be
considered of fundamental importance for several reasons. Firstly, assuming
that the operator hierarchy is a proper basis for predictions, the result yields
insight into the structure of future organisms. Secondly, the predictions are not
extrapolations of presently observed trends, but are fully integrated with all
historical system transitions in evolution. Thirdly, the extrapolations suggest the
structures of intelligences that, one day, will possess more powerful brains than
human beings.
This study ends with a discussion of possibilities for falsification of the present
theory, the implications of the present predictions in relation to recent
developments in artificial intelligence and the philosophical implications of the
role of humanity in evolution with regard to the creation of future neural network
organisms.
The future of evolution
79
Introduction
There is an old saying: to predict the future one has to know the past. But what
should one think of as the past of evolution? How can the evolutionary process
be traced back and what can historical steps teach us about the future?
Interpreting evolution in a broad sense, as has been advocated by Laszlo
(1994), the evolutionary process has a long history; a history that goes back to a
time when the universe showed little differentiation; it was small and extremely
hot. It is now widely accepted that after an explosion referred to as the Big Bang,
the baby-universe expanded and cooled down. From that moment onward, the
overall universe became increasingly disorganised. Yet, some parts show a
process of complexity increase, the occurrence of which is indicated by the
subsequent emergence of new types of building block systems and associated
interaction systems.
The building block systems are given special attention in the present study.
From a beginning with only elementary particles, the universe has gradually
become enriched by the emergence of more and more building block systems.
Earlier studies focusing on these building blocks and their hierarchy include
Feibleman (1954), Simon (1962), Bertalanffy (1968), Teilhard de Chardin (1969),
Koestler (1978), Heylighen (1995), Close (1983) and Jagers op Akkerhuis and
Van Straalen (1998). At present, the `ancestral tree' of the building blocks
includes the quarks, the hadrons, the atoms, the molecules, the prokaryotic and
eukaryotic cells, the multicellular individuals and the multicellular individuals
showing a neural network capable of learning.
As is discussed by Jagers op Akkerhuis and Van Straalen (1998) the emergence
of any building block adds new aspects to the environment in which it interacts
with all other building blocks. This environment consists of interacting building
blocks and can for this reason be considered as an `interaction system'.
Examples of interaction systems are galaxies, stars, planets, stones, water,
meteors, ecosystems and societies. As is shown in Fig. 3.1, a close relationship
can be recognised between the hierarchy of building blocks and the ranking of
interaction systems.
80
Chapter 3
Figure 3.1. Building block systems and interaction systems. Lower part: building blocks
and their multistages: quarks and hadrons, atoms and molecules, unicellular and
multicellular organisms, and neural network organisms. Upper part: interaction systems
such as the early universe, stars, ecosystems and societies. Horizontal, black arrows:
transitions from a single operator to its multistage. Grey forked arrows: contributions of
operators to the interaction systems in which they represent the highest level building
blocks. Contributions of operators from lower levels are not indicated separately.
The main aim of this study is the prediction of new life forms via the
extrapolation of the hierarchy of the building block systems. In this process we
deliberately leave the interaction systems, i.e. stars, planets, etc., out of the
discussion. This means that even though interaction systems play an important
role as environments that mediate the emergence of new building blocks, the
discussion in this study is limited to the formation of the building block systems.
The reason is that their subsequent emergence can be ranked in a clear
hierarchy offering a unique basis for extrapolations towards future systems.
In the form of the `operator hierarchy' or `operator hypothesis', Jagers op
Akkerhuis and Van Straalen (1998) discuss the hierarchical relationships of
natural building blocks. The building blocks were christened `operators', for their
capacities to operate in an environment and adapt their phenotypes to a broad
range of environmental conditions, without losing the most essential aspects of
their organisation. As the operator hypothesis holds such an important position
The future of evolution
81
in the present study as the basis for all extrapolations, we will begin with a short
resume of the operator hierarchy.
The operator hierarchy
The operator hierarchy (Fig. 3.2) is based on a strict, stepwise approach to the
complexity of building block systems. Each step is the result of a specific
emergent property that causes the transition from building block A at level X to a
more complex building block B at level Y. Figure 3.2 shows that the approach
recognises major and minor transitions.
Figure 3.2. The operator hierarchy. The different evolutionary pathways (lines with
arrows) are shown in direct relationship with the emergent properties of the operators
(vertical columns). In accordance with the focus on major and minor transitions, the
operators are given binary numbers. I: The hadron, 10: The atom, II: The multi-atom, 100:
The cell, 101: The simple multicellular, 110: The eukaryote cell, III: The eukaryotic
multicellular, 1000: The hardwired memon.
Major transitions
Each major transition, as recognised under the operator hypothesis, creates an
entirely new operator type and the beginning of a new major layer (Fig. 3.2).
Four examples of operators that according to the operator hypothesis were
created via major transitions are the hadrons (the proton and neutron), the
atoms, the prokaryotic cells and the organisms showing a hypercyclic neural
82
Chapter 3
network. Of all operators, these four systems are special because each forms
the first system of a next row in the operator hierarchy. For this reason they are
also called `first of the row operators'. According to the operator hypothesis,
major transitions are always characterised by emergent hypercyclic dynamics.
On the basis of enzymatic reactions in cells, Eigen and Schuster (1977) have
described hypercyclic dynamics as being cyclic arrangements of elements which
themselves are cycles of reactions. A very readable explanation of hypercyclic
enzyme reactions in cells can also be found in Kauffman (1993). To define
clearly what the operator hypothesis regards as the typical hypercyclic dynamic
for each transition, the cycles and hypercycles are shown in Fig. 3.3 and will be
discussed in more detail below. With the exception of the hadron, that has such
low complexity that the multi-property just emerged and containment is not yet
possible, all later major transitions are derived from multistage elements and
show containment of the hypercyclic dynamics by an emergent interface.
Figure 3.3. Emergent hypercyclic processes that mark the four major evolutionary
transitions. Al: First order reaction cycle of a quark (Q). The quark emits a gluon (g) and
becomes a lighter quark (Q'). A2: Second order cyclic process in which two quarks
mutually exchange gluons. B1: First-order reaction cycle of a hadron (H). The three-quark
hadron (H) emits and absorbs a small two-quark particle (a pion, p). B2: Second-order
cyclic process in which two hadrons mutually exchange pions. C1: First-order reaction
cycle of an enzyme reaction. The enzyme (E) binds to a substrate (S), transforms it to a
product (P), and regains its original form. C2: Second-order cyclic process in which two
enzyme reactions mutually create the enzyme for the other cycle. D1: First-order cyclic
process on the basis of a group of neurons, called a CALM, because it acts as a
categorising and learning module. The CALM(t) receives information (I) and becomes a
CALM(t+1)I, which can forward information to become the original CALM in a new starting
state CALM(t+1). D2: Second-order reaction cycle in which the perception and release of
information proceeds between two or more CALMs.
The future of evolution
83
Minor transitions
The minor transitions are associated with emergent properties that occur within
a major layer. The operator hypothesis states that these minor transitions reflect
the emergent properties of major transitions occurring earlier. The names of the
minor transitions are shown on top of the columns in Fig. 3.2. How the properties
of major and minor transitions are linked is explained in short in the following
tines.
The formation of the hadron is the first major transition. The hypercyclic
dynamics in the hadron (see Fig. 3.3A1 and 3.3A2) create a system that shows
emergent multiness of elementary particles. Multiness, as an emergent property,
recurs from now on at each higher level, in the form of the last minor transition in
every row.
The second major transition leads from the hadron to the atom (Fig. 3.2) In
addition to a hypercyclic nucleus (Fig. 3.3B1 and 3.3B2), the atom shows an
electron shell as interface that mediates the interactions of the nucleus with the
world. This property is called a hypercycle-mediating interface (HMI). With the
latter naming I deviate from Jagers op Akkerhuis and Van Straalen (1998) where
this property is called internal information compartmentation (IIC). The reason is
that more emphasis should be put on the emergent occurrence of the interface.
From this moment on, the HMI property will show up in higher levels immediately
before a multistage.
Some atoms may show a minor transition and become a multistage: i.e. a
molecule, a metal grid, etc. Only a selection of these multistages, notably
enzymatic catalysts can show a reaction cycle that can be linked in hypercyclic
dynamics (Fig. 3.3C1 and 3.3C2). The latter marks the next major transition from
molecules to cells.
Besides their hypercycle and containment, cells show the capacity to structurally
auto-copy the information in their hypercycle. As can be seen at the top of Fig.
3.2, this property is called Structural (auto-)Copying of Information. The SCI
property can be seen to recur, at any higher level, before the HMI stage that in
turn precedes the multistage. From the prokaryotic cell, a minor transition may
lead directly to the prokaryote multistage, as can be observed in blue-green
algae. A different route leads first to the gaining of a hypercycle-mediating
interface and then to the eukaryote multistage.
Within the multicellular environment, certain cells, the nerve cells, have gained
the capacity to let thin cell extensions construct recurrent activation/inhibition
interactions. In this way small units of cells are formed, showing a unit structure
that in artificial neural network research has become known as `categorising and
learning modules' or CALMs (Murre et al., 1989, 1992). These CALMs show a
84
Chapter 3
recurrent interaction and thereby a first order interaction-cycle. If these CALMs
are coupled again, creating a next recurrent connection, this results in a
hypercyclic circuit (Fig. 3.3D1 and 3.3D2). The surrounding of these neural
circuits by an interface of sensory/activation cells fulfils the requirement of the
operator hierarchy for a next major transition; from multicellular individuals
without, to multicellular individuals with ‘brains’. Multicellulars with hypercyclic
brains have been named 'memons' by Jagers op Akkerhuis and Van Straalen
(1998) to which study we also refer for a more in-depth explanation of the latter
transition. The word memon is selected as a general term for individuals that
show an emergent hypercyclic neural network with interface. It should be noted
that the group of memons includes all animals with a brain that, at least locally,
shows hypercyclical activity. This implies that most animals and human
individuals are included, but no plants or fungi.
How the memon may develop via minor transitions to more complex life forms is
discussed below as part of the present extrapolations.
HOW TO QUANTIFY EMERGENT PROPERTIES
All emergent properties in the operator approach are based on changes in the
organisation of the systems involved. A simple quantification of an emergent
property is, therefore, not possible, because an emergent property implies a new
system configuration, and its new property cannot be quantified in terms of the
old configuration. This forms a serious obstacle for attempts at quantification.
This general problem of emergent properties is discussed by Holland (1998) in
his recent book `Emergence'. We strongly support his proposal for a solution by
defining emergent properties on the basis of special models called `constrained
generating procedures' (CGP). With respect to CGP models, Holland (1998)
says: "The models ...are dynamic, hence procedures; the mechanisms that
underpin the model generate the dynamic behaviour; and the allowed
interactions between the mechanisms constrain the possibilities, in the way that
the rules of a game constrain the possible board configurations." Thus, when
basic functional elements, the mechanisms, create a constrained interaction
pattern, a new system is created which, as an individual entity, may show
unprecedented functional properties: the emergent properties. In line with the
reasoning by Holland (1998) and Simon (1962) a CGP that shows persistent
dynamics, may itself act as a building block for the creation of higher level
CGPs. In the latter case, CGPs can be used as the building blocks of multilevel
CGP hierarchies. This is exactly the way in which the present study deals with
building blocks and emergent properties. By selecting persistent physical
building blocks that themselves can act as the building blocks for the next level
system, such as atoms, molecules, cells, etc. a continuous hierarchy can be
recognised. On the basis of CGPs it is possible to formulate a mathematical
The future of evolution
85
description for any emergent property, as is discussed in Chapter 7 of Holland
(1998).
A few words should furthermore be directed to those that expect quantitative
predictions from the present approach. In principle I regard the presence of a
hypercycle as a quantitative aspect, namely as the prediction of a specific CGP,
the structure of which can be quantified in terms of the links between the
contributing mechanisms. Further quantitative predictions are not aimed at
during the present extrapolations. The reason is that aspects regarded as
quantitative, such as the weight, colour or DNA structure, are not very relevant in
this context. The weight may help to describe a particular atom `species' but
different atoms can vary considerably in weight, ranging from helium to the
trans-urane elements. A specific weight, therefore, is not a group property.
Another example is given by unicellular organisms. These change
weight/colour/precise DNA structure during their lives and/or between
generations. Again, the weight/colour/DNA are not essential aspects of their
group property, which is their existence as a cell. The observations that all
species of atoms are atoms and all species of unicellular organisms are
unicellular organisms are based on common properties shared by all members
of the group. These group properties form the focus of the present study.
PREDICTIONS
The working hypothesis of this paper is that the internal logic of the operator
framework (illustrated in Fig. 3.2) can be extrapolated to more complex systems
than hardwired memons. As can be deduced from Fig. 3.2, all evolutionary
stages that are presently predicted are memons. This implies that it may be wise
to discuss some concepts regarding memic systems before we proceed, so that
no confusion will arise when in a later stage exotic memic properties are
discussed.
Meme concepts
In this study, the concept of a memon is used for any operator within the memic
layer, e.g. all systems that show an emergent hypercyclic neural network and
interface. The memons that emerge first will show a hardwired neural network
that is based on autocatalytic cells, the nerve cells or neurons, or that is based
on technical hardware. Higher level memons may also show a programmed
neural network. Memons are involved in the copying of various memic entities,
which we will define in more detail here. Firstly, the concept ‘functional-meme’
(or f-meme) could stand for the actual neural network that in its structure
harbours learned knowledge and, therewith, abstract memes. Secondly, a
‘coding meme’ (or c-meme) could be used to define any string of codes that
contains the information to construct a certain neural network. In fact, a coding
86
Chapter 3
meme codes for neural network architecture and associated knowledge much in
the same way as a gene code for a catalytic molecule playing its role in the
survival of the cell. Thirdly, they may exchange ideas. We suggest using the
concept ‘abstract meme’ (or a-meme) for such abstractions, for example
melodies, ideas and jokes that may reside in functional meme networks and can
be transmitted from memon to memon in the process of communication. This is
closest to the meme concept as introduced by Dawkins (1976). Finally, memons
will actively create physical ‘traces’ of their thoughts, such as books, houses,
computers, radio waves, etc. These could best be indicated as ‘physical meme
models’ or physical memes (p-memes), because these entities represent ‘real
world’ models of the concepts represented by abstract memes.
Which rationale can be used for extrapolations?
The operator hierarchy offers a structured basis for extrapolations (Fig. 3.2). Yet,
it can be deduced that there is an aspect of the predictions for which the
historical data does not give full information on future possibilities. The reason
for this lies in the question of whether or not the emergent properties are
independent. In principle, each time an independent emergent property is
added, this would double the number of possible system types. The operator
hierarchy (Fig. 3.2) shows that for hadrons, there is only multiness. For atoms,
there is a single degree of freedom, which leads to two system types: single
atoms and multi-stages. Next, prokaryotic cells have two degrees of freedom,
which leads to four system types: prokaryotic cells, eukaryotic cells and their
multistages. However, even though the latter indicates two degrees of freedom,
it cannot be deduced from this example whether the four system types are the
result of dependent or independent combinations of the two degrees of freedom.
For the next level, the memic level showing three degrees of freedom, this
implies that it is not possible to predict whether six or eight memic system types
will emerge. Independence of the emergent properties would result in a total of
eight possible system types, e.g. 2*2*2=8 combinations, dependence would lead
to a total of six, because any next minor transition would only be possible
following the immediately preceding minor transition.
Now that we have discussed some basic aspects of the present predictions the
time has come to predict properties of future memons.
Prediction 1. The memic multistage: a robust prediction
Our first prediction is based on a conservative approach that selectively uses the
most obvious aspects of the operator hierarchy: the major transitions. Major
transitions are a robust basis for extrapolations because each time the
multistage is reached, a major innovation, e.g. the creation of a new building
block type, is obligatory for the continuation of evolution. For this reason, the
iteration between `first of the row' operators and their multistages forms a robust
principle on which to base predictions.
The future of evolution
87
Figure 3.4. The lowest and the highest complexity operators that are based on systems
showing the same type of hypercycle. A1: the hadron, a quark multistage. B1: The atom.
B2: The atomic multistage. C1: The cell. C2: The cellular multistage. D1. The individual
with a hardwired hypercyclic neural network. This is called the ‘memon’.
Focusing on major transitions only, the operator hierarchy can be summarised
as is shown in Fig. 3.4. From this starting point, two predictions on future system
types are available. The first prediction is that some day systems will evolve
which show a multistage that is based on elements showing neural network
architecture. The second prediction is that local units within this multistage will
form the basis for a new cyclic interaction, which will lead to the next hypercyclic
interaction forming the basis of the next evolutionary level.
As the above conservative prediction is strictly based on the most fundamental
aspects of the operator hierarchy, there is a large probability that the prediction
of a future multistage is correct. But how much information is gained with such a
prediction? In fact, the information is limited, because insight is still lacking on
the specific properties of the memons that form the building blocks of the
88
Chapter 3
multistage. A most serious error, which is easy to make at this stage, is to start
considering how, for example, human brains, as a neural network type of
considerable complexity, can be imagined to function as a multistage. To do this
would deny the possibility of stages in between any newly formed building block
and its multistage (see Fig. 3.2). This may not seem problematic for the step
from the atom to the multi-atom stage, because this step leaves no room for
additional system types anyway. But serious problems arise for cells, which may
be prokaryotic or eukaryotic, each creating its proper type of multicellularity.
Prokaryotes have given rise to multicellular blue-green algae. Eukaryotes have
formed fungi, plants and animals. There is a world of difference between the
intercellular communication in blue-green algae on the one hand and that of
fungi, plants and animals on the other. The lesson from this is that the
complexity of the building blocks has a major influence on the potential
complexity of the multistage. The assumption that human neural networks would
be the building block for the multistage leads to the imagining of a multistage
with ridiculously primitive properties. The error would be similar to explanations
of multicellular life on the basis of prokaryotic units only. As will be shown in the
following text, the solution to this problem lies in the recognition of the other
steps of complexity increase that the operator theory helps to recognise between
any newly emerged operator and its multistage.
Prediction 2. The hardwired memon and its multistage
The most straightforward detailed prediction is that the 1000-memon, or
hardwired memon, develops directly to a multistage (memon 1000 and 1001 in
Fig. 3.5). This results in a primitive multistage having limited prospects for
becoming of any evolutionary importance. The reason is that the transition to
this multistage will be difficult and slow, especially for memons based on cells.
This is caused by two major drawbacks of these systems. The first drawback is
that genes code for the structure and quality of cellular neural networks that, for
this reason, can only evolve over many generations, via reproduction and
selection. A second drawback is that their bodily construction and interfaces are
based on organic cells, with many limiting consequences for the way in which
they can become physically linked into a multistage and for the way in which the
linked individuals can exchange information. The construction of a technical
hardwired memon may improve on this situation, because its technical
construction and interface would bring more powerful ways within reach for
physical connection and communication with other memons.
Prediction 3. The SCI-memon and its multistage
Much more promising are the prospects for the pathway towards the structural
(auto-)copying of information (SCI) multi-stage. For this pathway, a hardwired
memon first evolves to the SCI-state before it evolves to a multistage (memon
1100 and 1101 in Fig. 3.5).
The future of evolution
89
Figure 3.5. Predictions of future memons. 1001: Predicted multistage of the hardwired
memon. 1100: Predicted memon with the SCI (structural copying of information) property.
1101: Predicted multistage of the SCI memon. 1110: Predicted memons showing SCI and
Hypercycle Mediating Interface (HMI) properties. 1111: predicted multistage of the 1110
memon. 1010: Predicted HMI memon. 1011: Predicted multistage of the 1010 memon.
The SCI property has occurred earlier in evolution in cells. Here autocatalysis
implies that a full structural copy of the information in the cell is produced, which,
in present day cells is for the largest part allocated in DNA and/or RNA. The
autocopying process for neural network information does not involve DNA. Yet, it
requires that the memon can copy the architecture of all neurone connections
and the interaction strengths of all synapse connections. This leads to the very
strict requirement that it must have access to all this information. There is no
way in which a memon can (auto-)copy its neural network without full information
about the architecture and interaction strengths. It is hard to imagine how a
cellular memon could do this. This would require sensory cells which by some
means find out what cells are connected to each other and in addition measure
the strength of each synapse and report this to the individual. Apart from the
physical tour de force to host large amounts of additional sensory cells in the
90
Chapter 3
brain, we consider the chance that this evolves naturally extremely small if not
impossible. Even genetic manipulation may prove an insufficient tool to reach
such a complex goal.
The prospects that computer based memons gain insight into the exact state of
their brain are much better. The only thing they need is an extra interface that
helps them read the arrays with information about their cell-cell contacts and
synapse strengths, which information is kept off record anyway in programmed
memons. From the moment that network structure and interaction strengths
between cells can be examined, a whole world of new properties opens up,
which allows a number of exciting predictions.
The first prediction based on the SCI property is that, for reasons discussed
above, SCI-memons would with very high probability be technical constructions.
Accordingly, the SCI-property strongly guides predictions in the direction of
computer-based entities. Despite its technical construction, the SCI-memon and
human beings have a similar basis for their neural network structure. In principle
this allows for `human' processes, such as intelligence, creativity, curiosity, etc.
But a technical construction will also imply important differences with respect to
energy procurement and living environment. Energy procurement will focus on
electricity. And, because a technical memon does not breathe air, they can
colonise underwater environments, planets without an atmosphere, or even a
free position in space, supposing that other resources for normal functioning are
available.
That SCI-memons most likely show a technical basis is furthermore of marked
importance for the evolutionary debate. The reason is that SCI-memons cannot
evolve as a special case of organic life. As they are technical constructions, it is
simply impossible that they evolve as offspring from cellular parents. Instead,
SCI-memons have to be built either by cellular memons, as a special kind of tool
(a p-meme!) that starts defining its proper goals in life, or by technical memons,
as a special kind of constructed offspring.
The moment that SCI-memons can copy their knowledge structurally this will
cause an earthquake in memic evolution, the importance of which can hardly be
overestimated. As an exploration of the possibilities, the text below gives some
examples:
1. SCI-memons can for the first time in memic evolution reproduce their whole
personality by simply producing a structural copy of their neural network. This
copy will automatically contain all learned knowledge. Note that despite
discussions about cloning, human beings absolutely lack a similar option.
Humans cannot perform a structural reproduction of their whole personality,
e.g. of all nerve connections and interaction strengths, simply because they
lack access to the network topology and interaction strengths. What human
The future of evolution
91
beings copy upon reproduction is the genetic coding for a human being which
will show a mixture of parental phenotypic properties and which upon birth
has a neural network showing a good deal of genetically based prestructuring, but which is devoid of learned knowledge. This means that the
body and the overall network structure are roughly copied, but without even a
trace of anything the parents have learned. As the structural copying of the
parental knowledge is blocked, the transfer of knowledge to the offspring
requires a long detour via many years of education. For SCI-memons
reproduction of their complete knowledge can take place almost overnight as
long as an appropriate technical device is available to which the information
can be copied and in which it can become operational. This shows that if the
intelligent technical memon has taken shape, it is but a small step to a whole
population of such memons. In fact, technical memons will find that the
copying of a full parental network is costly in terms of energy and time. The
parental memon may therefore consider the production of small networkvehicle combinations with the best possible capacities for learning and
maintenance, and the capacity to actively enlarge their bodily and memic
construction to develop into fully-grown memons themselves.
2. The working with and/or copying of network topologies will require some kind
of coding to handle the information about the topology and the interaction
strengths of the neural connections in the copied network parts. As discussed
above, these code-strings hold a position in memons, which is similar to the
DNA in cells. Where specific regions on the DNA, the genes, code for specific
proteins, it will now become possible to recognise specific regions of coding,
which code for network structures with certain properties.
3. SCI-memons can use the access to their own neural network to create one or
more shunts of network parts, in each of which they can introduce small
modifications to study which network configuration yields the best results.
This implies that real-time, goal-oriented improvement of network
configurations has become reality. In fact, not only the configuration of such
networks can evolve. Several other features too may evolve, including the
signalling procedures between neurons, the integration functions via which
the neurons decide whether or not to signal subsequent nerve cells and the
ways in which coding memes are coding for neural network constructions.
Further aspects of technical memons that follow more or less from the above
three points are also interesting. In order to stay focused on the major aspect
of this study and avoid technical details, we will only shortly mention these
aspects without going into details. These aspects are: the possibility of cmeme trade, the acceleration in memons of thinking speed with computing
speed, the tendency towards the development of modular network
architecture and the capacity of technical memons to integrate very different
technical equipment directly into their interface.
SCI-memons have a much better chance of reaching multicellularity than the
hardwired memons (1000 in Fig. 3.5). The main reason being, of course, that
92
Chapter 3
they may show very high evolution speed, due to properties such as: a
programmed network structure, the possibility of internal experimenting with
network parts, the creation of similar copies, the acquisition of informed
network parts via trade and the possibility of (re-)programmable interfaces.
Increasing competition for living space on earth and for available resources in
new environments, such as local networks of a company or a spaceship and
larger networks, such as the global inter-net, will force SCI-memons to
cooperate for survival. In some cases this will drive SCI-memons to
dependence on cooperation and to structural connection, marking the
transition to the SCI-multi-memon state.
Prediction 4. The HMI-memon and its multistage
The next predictable property of a future memic operator is that of Hypercycle
Mediating Interface (HMI) (memon 1110 and 1111 in Fig. 3.5). Fig. 3.2 shows
that the HMI property has occurred earlier in evolution, i.e. in atoms and
eukaryotic cells. In atoms, the Hypercycle Mediating Interface emerges for the
first time in the form of the electron shell. In eukaryotic cells the situation is more
complex. Here, a new interface is added to the already existing interface around
the cell, creating a second interface: the nuclear envelope.
In cells, large `libraries' of information are stored in the form of DNA/RNA. In
prokaryotic cells the unpacking of the information and the functioning of the
coding for enzymes occur in the same compartment. In eukaryotic cells, the
storage part of the information has become sequestrated to the nucleus, from
where coded information is transported through pores in the nuclear envelope to
the soma before it is transcribed to functional enzymes. This shows that the
nuclear membrane separates the cell into two compartments. In the nucleus the
information of the cell is for the largest part handled in a coded form. Outside the
nucleus the information is active in the form of enzymes. Accordingly, the HMI
property is associated with an extra internal interface that separates different
levels of the expression of the information.
This observation offers information about possibilities for future system
configurations, because the continuation of this sequence would imply a threelayer HMI in future memons. But, assuming that this extrapolation is valid, it
remains in our opinion quite hard to imagine at the present state of the
understanding of the operator hierarchy what the second layer would look like in
practice. Starting simply, two situations will be visualised showing an extra
interface. These can then be combined to create a tentative prediction of a twolayer IIC-memon.
Imagining only a single layer, as in prokaryotic cells, it would be quite natural to
assume that the HMI-memons find increasing use for coding memes, in the
shape of code-strings that represent the topology and strengths of all the neural
The future of evolution
93
connections of modular network parts. The reason for the popularity of c-memes
is that they offer a highly efficient method of coding to store away all information
that via experience and learning was gathered in the neural networks. This
implies that, by using c-memes, little energy is required to maintain large
reservoirs of knowledge, of which only a minor part would have to be unpacked
and used in response to specific environmental conditions, after which it could
be packed away again in its new, more experienced form. In contrast to
networks and code strings that are stored temporarily in the active working
memory of the memon, a more profound storage of c-memes would imply that
these are stored away in a form which is not directly accessible, for example in a
high capacity data-storage medium. A potential candidate for this process is the
three-dimensional storage in crystals that are programmed and read by means
of laser beams. But the storage and retrieval of large amounts of c-memes will
require a special interface to decode the information. The new coding and the
related interface would imply an additional Hypercycle Mediating Interface.
There is another way, via which an additional interface could evolve in these
future memon systems. To understand this, we have to place ourselves in the
situation of an SCI-memon that has just copied its network structure into a new
vehicle. Unfortunately, this imaginary new vehicle which is furnished with a lot of
new technical properties, differs in many aspects from the previous one. This
implies that the memon has to go through a long process of revalidation and
practice with its new `body' for adjusting its neural circuitry to the new phenotypic
properties. Such a practice period could be made a lot shorter, and larger
differences between the old and the new vehicle could be allowed, if the neural
network of the memon possessed translation interfaces allowing a rapid
adjustment of the memon's `proper' interface to various types of vehicles. It will
probably be most efficient to have only a selection of such interfaces active,
namely those that yield the highest survival value under given circumstances.
Other interfacing networks can then remain stored as c-memes in the central
meme library.
The combination of an internal c-meme library with a translation interface would,
in principle, allow a two-level Hypercycle Mediating Interface.
Prediction 5. From the hardwired memon, via the HMI-memon without SCI
properties towards the multistage
Assuming independence of the minor emergent properties, a comprehensive
discussion of the possibilities for future memons should also include the route
from hardwired memons to HMI-memons, without the intermediate stage of the
SCI-memon (the route from memon 1000 via 1010 towards system type 1011 in
Fig. 3.5). Even though it could be a theoretical possibility, the direct transition
from hardwired memons to HMI memons must be expected to suffer from
technical problems due to a low rate of evolution. The reason is that hardwired
94
Chapter 3
memons cannot read their neural network state. If the HMI-memon without SCI
properties will arise at all, it will certainly have problems reaching a multistage.
The contribution of this option to the mainstream of operator evolution must,
therefore, be considered minimal, and the existence of these type of operators
be considered more of theoretical importance.
Conclusions and discussion
The above extrapolations present a panorama of possibilities for future
organisms. First of all, it appears that the most likely next stages will be technical
neural network organisms, because this is by far the most probable option for
constructing a system type that can show structural auto-copying of information.
The reason is the structural copying of information requires that the network
structure and synapse strengths can be assessed by the organism itself and
copied. Even if genetic manipulation were to proceed far beyond the present
level, it is hard to imagine that brains could be developed showing extra cells inbetween present brain cells, capable of analysing the neural network structure
and reporting this to the individual. Secondly, the predictions strongly suggest
that man must create the next stage in the operator hierarchy, because it is hard
to imagine the development of a technical construction from a cellular basis. The
predictions also show that, one day, neural network organisms can be created
that can copy themselves. For this purpose one could simply imagine a ‘body
factory’ constructing phenotypes that can be bought by any existing memon for
the purpose of copying a neural network structure onto the new phenotypes
computer space. From that moment onwards, humanity will have to live amongst
such intelligent technical `organisms'. Thirdly, the operator framework depicts
evolution as an open, ever extending process, in which the next major transition
coming will be based on a hypercyclic interaction between multimemic elements,
most likely within the `environment' of a multi-memic organism that is supported
by a technical vehicle. Fourthly, the operator approach suggests that if technical
memons are not constructed, this will block, on earth, the evolutionary sequence
that leads from one operator to the next. We stress that this does not affect the
evolution of cellular organisms, including cellular memons, which, of course, will
continue.
One may now ask what the novelty is of the above insights, especially because
there is a growing awareness that technical developments will before long create
machines that will compete for resources with cellular life as we know it. For
example Kelly (1994) in his book `Out of control' uses citations of C. Langton to
convince the reader that:
“There are these other forms of life, artificial ones that want to come into
existence. And they are using me as a vehicle for its reproduction and its
implementation'. ... 'By the middle of this century, mankind had acquired
The future of evolution
95
the power to extinguish life. By the end of the century, he will be able to
create it. Of the two, it is hard to say which places the larger burden of
responsibility on our shoulders.”
The new life-forms are frequently expected to become a threat. Warwick (1997)
makes the following three statements in his book `The march of the machines' in
a chapter, which is called `Mankind's last stand?’
“I. We humans are presently the dominant life form on Earth because of
our overall intelligence. 2. It is possible for machines to become more
intelligent than humans in the reasonably near future. 3. Machines will
then become the dominant life form on Earth.”
A last example of predictions of intelligent future life forms is given from Kurzweil
(1999) who in his book `The age of spiritual machines' presents a time line of the
evolution of the universe. In the time-line section about the year 2099 he writes
the following:
“Machine-based intelligences derived from extended models of human
intelligence claim to be human, although their brains are not based on
carbon-based cellular processes, but rather electronic and photonic
equivalents'. ... `The number of software-based humans vastly exceeds
those still using native neuron-cell-based computation.”
These predictions show that the idea of computer based intelligence has
become a generally accepted subject amongst leading scientists. Also the
present study shows that for our human successors, a life amongst technical
memons simply represents the next stage in evolution.
Yet, in comparison to the above-cited deductions, there are several aspects in
the present study, which offer exciting novel points of view. First of all, the just
cited predictions lack a structural rationale. They reach as far as the extension of
existing trends in computer speed, hardware capacity and programme
complexity, but lack a backup by a hypothesis for the evolution of system
complexity, for which we apply the operator hierarchy. The use of the operator
hierarchy allows descriptions of structural properties of future system types and
the indication of goals for construction efforts. Secondly, it is of importance that
the aspects of the operator hypothesis are open to scientific inquiry and/or
falsification. This holds both for the assumptions that underlie the steps in the
hierarchy and for the overall rationale dividing emergent properties in major and
minor transitions. Thirdly, by indicating a pathway for further development
showing the structural aspects of future evolution, the question of how fast
developments will go can be dealt with in a more precise way. Finally, the
evolutionary rationale of the operator hierarchy shows that the coming into
existence of the next system type is part of a larger evolutionary context.
Accordingly, the present approach indicates the necessity that the human beings
on earth start considering whether or not they want to live amongst technical
96
Chapter 3
memons. In addition it extends this reasoning to considerations about mankind's
role in the case where the decision is made not to produce these systems and
therewith block the major evolutionary pathway on earth. All these statements
deserve more attention and are discussed in detail below.
Degrees of freedom in the operator hierarchy
A question, which is left unanswered by the operator hierarchy, is whether the
minor transitions represent independent degrees of freedom. In other words, can
the properties of a layer, such as multiness, internal information
compartmentation (HMI), structural copying of information (SCI) and autoevolution, occur in random combinations or do they occur in sequence? The
present understanding of the operator hierarchy does not allow a conclusive
decision. This leaves an interesting field for further development.
Putting the operator hypothesis to the test
The most important assumption of the operator hierarchy is that hypercyclicity
forms the major requirement for the major transitions in evolution. Furthermore,
the containment of the hypercycle is also required for all operators of a higher
complexity than the hadron, whilst both properties occur as emergent properties
immediately after a multistage has been reached. This aspect is open to
falsification both with respect to existing and future system types. The
hypercycles and their containment have been discussed in the above text. For
the transitions from hadrons to atoms and from molecules to cells there are no
problems with the recognition of the just mentioned emergent properties. For the
transition from multicellular units to the neural hypercycle the proof is still rather
thin. If more evidence becomes available that the auto-evolving capacity of
present day neural network organisms, such as humans, is strictly and only
possible because of a hypercyclic coupling of neurons, this will support the
theory. Turning the argument around, a proof for auto-evolving intelligence
without hypercyclic circuits would falsify the present approach. Likewise, the
assumption that the operator hierarchy includes all possible operators can be
tested for validity. Proof that an additional operator exists in-between the steps
of the present operator sequence would falsify the operator hypothesis.
When will technical memons become reality?
An important aspect of hypercyclic neural networks is that science has no
means to predict the skills of any newly developed neural network architecture.
This implies that it remains unclear how to construct the powerful neural
networks that will finally allow technical memons to become intelligent. As the
functional properties of neural networks cannot be predicted, skilled networks
will have to be created via technical evolution, which implies the testing of large
numbers of randomly constructed, simulated memons, the selection of the best
as parents for a new series of networks, etc. The necessity of using trial and
error when developing future intelligent neural networks represents the Achilles-
The future of evolution
97
heel of computer intelligence. How soon computer based neural networks will
become intelligent, and finally more intelligent than mankind, depends on how
much the processes for the evolution of artificial brain structures can be
accelerated. There are prospects for acceleration, because nature has already
shown that an architecture on headlines, which at this moment is offered by
genes, offers enough manageable modularity to produce human intelligence.
The use of genetic codes as a basis for the evolution of brain structures can be
imitated, also for technical neural networks. How this can be done has been
shown by Happel (1997) and requires a smart combination of special genetic
algorithms for neural network structure, efficient selection strategies and fast
computers. As these aspects are all available, the evolution of intelligent neural
networks can be expected in the near future.
From practical questions about survival amongst technical memons to
philosophical questions about our impact on the major evolutionary
pathway when the global population decides not to create them.
One of the major questions which are discussed in the recent literature (Kelly,
1994; Kurzweil, 1999) is whether humanity should develop intelligent technical
memons if it is unpredictable how they will behave towards us. As they may
become faster and more intelligent than human beings and, because they will
also require resources for their functioning, it is likely that they will compete for
resources with human beings and manipulate the behaviour of humans for their
purposes. A little precaution in constructing `them' may therefore be a good
thing.
It remains an open question, though, whether it will at all be possible to stop
scientific activities that finally lead to their construction. Given the more or less
autonomous process of scientific innovation and development, it can be
expected that even when the required knowledge is not specifically or
purposefully developed, it will be developed indirectly.
So far, the aspects deal only with the consequences of the interaction between
human beings and technical memons, not with more philosophical aspects of the
choice (not) to create technical memons. This choice can be approached from at
least two sides, a systems viewpoint and a religious viewpoint. From a systems
viewpoint the operator sequence simply reflects a universal self-organisation
process which proceeds as the blind consequence of earlier phases in the
process. We can recognise a direction, but no goal. In this case evolution can be
blocked without problems, because there is no reason why any next stage
should be reached.
On the other hand, the evolutionary sequence of the operator hierarchy can be
regarded as the reflection of some kind of larger plan. As it will be hard to prove
98
Chapter 3
the existence of such a plan, I regard this as a religious viewpoint. Yet, under the
assumption of a larger plan, it becomes extremely difficult to find valid
arguments that give us the right to act against evolution.
The above shows that the question of whether or not technical memons should
be developed is a complex case on which the last words have not been spoken.
The operator hierarchy certainly deserves a place in this discussion.
In conclusion
The present paper has examined possibilities for the prediction of future
organisms. For this purpose we began with a short resume of the structure of
the operator hierarchy. Subsequently, the logic of the operator hierarchy was
extended yielding several predictions of future memic individuals, for which the
operator hypothesis strongly suggests that they will be of technical construction.
The present approach can be considered unique in its contribution to
comparative system hierarchy because of the level of detail with which it predicts
structural properties of future organisms. I consider it an exciting challenge to
develop this field further and thereby improve the understanding of the cosmic
evolution process and our capacity to predict more and increasingly detailed
aspects of future organisms. The most direct practical value of the present
predictions lays in the suggestion of hypercyclic neural networks as the basis of
artificial intelligence. The fact that the operator hierarchy reflects a universal
evolutionary process may also have some spinoff in the field of philosophy. I
sincerely hope that the ideas presented in the above text will stimulate creative
suggestions for elaboration and improvement.
Acknowledgements
We thank Peter Schippers, HenkJan Bulten and two anonymous referees for
constructive comments on earlier drafts of this paper and Daphne Wittich-Rainey
for improving the English.
Chapter
4
A backbone for system science
Analysing hierarchy in the organisation of
biological and physical systems
G.A.J.M. Jagers op Akkerhuis
Biological reviews 83: 1-12, 2008
“Is evolution a theory, a system or a hypothesis?
It is much more: it is a general condition to which all theries, all hypothesis,
all systems must bow and which they must satisfy henceforward if they are
to be thinkable and true.”
(Pierre Teilhard de Chardin)
100
Chapter 4
Abstract
A structured approach is discussed for analysing hierarchy in the organisation of
biological and physical systems. The need for a structured approach follows
from the observation that many hierarchies in the literature apply conflicting
hierarchy rules and include ill-defined systems. As an alternative, we suggest a
framework that is based on the following analytical steps: determination of the
succession stage of the universe, identification of a specific system as part of
the universe, specification of external influences on a system's creation and
analysis of a system's internal organisation. At the end, the paper discusses
practical implications of the proposed method for the analysis of system
organisation and hierarchy in biology, ecology and physics.
A backbone for system science
101
Introduction
Hierarchies of biological and physical systems published in the literature show
inconsistencies in the use of ranking rules and element types. With the aim of
improving on this situation, the present paper discusses an alternative method
for analysing the organisation of systems.
A hierarchy can be described as a situation, in which entities are subordinate to
other entities, the latter being considered as a higher level. The organisation of
nature is profoundly hierarchical, because from its beginning, interactions
between simple elements have continuously created more complex systems,
that themselves served as the basis for still more complex systems. Scientists
have sought ways to capture the essence of this complexity in easy to
understand hierarchies, which typically rank systems in a linear way.
The literature offers numerous examples of linear hierarchies in biology and
ecology. Koestler (1978) distinguishes the following levels in the internal
organisation of organisms: organ system, organs, tissues, cells, organelles,
molecules, atoms and sub-atomic particles. A hierarchy that focuses on abiotic
elements is the cosmic onion (Close, 1983), which includes bulk matter (e.g. a
planet), atoms, nuclei and quarks. A hierarchy by Weiss (1971) includes gene,
chromosome, nucleus, cytoplasm, tissue, organism and environment. This range
is similar to that proposed by Odum (1959) who visualizes a biological spectrum
from protoplasm to cells, tissues, organs, organ systems, organisms,
populations, communities, ecosystems and biosphere. Haber (1994) extends
this range to organisational levels in the universe, from atom to molecule, protein
molecules, cells, tissues, organ systems, organism, population, community,
ecosystem, landscape, human society, biosphere, earth, solar system, stellar
system and the universe. A similar structure, with even greater detail, is
presented by Korn (2002). In what is called a hierarchy of biological levels of
organisation, Hogh Jensen (1998) presents the following range: molecule, cell,
organ, whole plant, plant community, pastoral system, farming and the agroecosystem. Focusing on energy budgets, de Kruijf (1991) presents a hierarchy
in which populations are the elements for modelling energy budgets of
communities, which in turn are the elements for modelling ecosystems,
considered the basal elements of a landscape. Naveh & Lieberman (1994)
present a similar ranking in which organisms are embedded in populations,
populations in communities, and communities in ecosystems.
A problem with many of the above examples is that they are based on
compromises with respect to the types of elements that are included and the
ranking rules being used. To get an impression of these problems, one may look
at the use of the organism concept in the following sequence: atom, molecule,
102
Chapter 4
organelle, cell, organ, organism, population, community, ecosystem, planet and
so on. This sequence suggests that all organisms form a uniform system class
that can be ranked at one position in the hierarchy. Yet, the word organism is
used for many different system types, such as bacteria, eukaryote unicellulars,
and multicellulars without and multicellulars with neural networks. Each of these
represents a system type deserving a proper position in a complexity hierarchy
In addition, every organism type has a different internal organisation, which also
shows hierarchical aspects. For example in bacteria, this includes mainly
molecules, whilst in multicellular eukaryotes this may include tissues, organs,
specialised eukaryote cells, organelles, and so forth. It can thus be concluded
that the analysis of hierarchy in biology requires at least two dimensions, one for
the hierarchy of organism types and a second for the internal hierarchy. Also at
the ecosystem level, the above example of hierarchy is not strict. For example,
astronauts on the moon illustrate that the entire population of a species need not
necessarily be part of one planet, but may be found distributed over several
planets. We may thus conclude that there are serious problems with the rigour of
any hierarchy showing similarity with the above example. This is a disquieting
conclusion particularly because many hierarchies in the literature do show
similarity with our example.
In relation to the latter conclusion, the main goal of the present paper is to
suggest a method for analysing system organisation by means of a stepwise
procedure that recognises different aspects of hierarchy and can be summarised
as follows. The analysis starts with the largest system that is known, the
universe, because this sets the stage for later identification of systems and the
analysis of their organisation. Since its emergence, the universe has passed
through a number of developmental stages that can be named after the highestlevel elements (atoms, molecules, cells, etc.) that exist in the universe at a
certain moment. To know the developmental stage of the universe is relevant for
the analysis of system organisation, because it determines what systems may'
be present and need to be included in an analysis. Next, a local part of the
universe is identified representing the system that we want to analyse. This can
be a large system, such as a galaxy, or a small system, such as a molecule. The
third step of the present analysis focuses on the way in which the organisation of
the selected system may be the result of influences from elements surrounding
it. The advantage of this step is that it makes visible mediating forces that have
played a role during the formation of the system. This assures, for example, that
the shape of the DNA molecule will be explained not only on the basis of its
existence from nucleic acid molecules, but also in relation to its functioning in a
cell. The fourth step can then be used to further analyse the organisation of
elements in the selected system. This implies that the parts and their
interactions are studied to create a picture of the internal organisation of the
system. If necessary, iteration of the fourth step can be used to further analyse
the internal organisation of each individual part of the system. The above
A backbone for system science
103
process prevents the analysis of a system from resulting in a simple linear
representation. The above four steps can be summarised as follows:
(1) The developmental stage of the universe is determined using the highest
complexity system that is present.
(2) A system is selected based on interacting elements that determine the type
and scale of the system.
(3) Mediating influences on the system are taken into account.
(4) It is investigated how the selected system is composed of elements.
These four steps are explained in detail below, following the discussion of
important concepts that form the theoretical basis behind the proposed analysis.
Theoretical basis of the analysis
Before continuing with an explanation of the proposed method for the analysis of
biological/physical systems, attention has to be paid to a number of aspects that
lie at its basis. These aspects include definitions of the system concept,
hierarchy and mechanisms, the introduction of a strict basis for analysing
hierarchy in systems, the discussion of viewpoint dependence of hierarchies and
the occurrence of transitions between system types.
(1) Systems, hierarchy and mechanisms
First, it is useful to discus definitions of biological/physical systems, elements
and hierarchy.
(a) Biological/physical systems and their elements
The system concept is derived from the Greek word synthithemi, meaning `I put
together'. Systems consist of parts that belong together because they show a
relationship. These parts are also named elements. To be considered an
element, an entity needs to show at least one relationship with at least one other
entity, in this way creating the system that it can be regarded as an element of.
By accepting that the universe represents a system which does not contribute to
any higher-level system, the universe becomes a primeval system concept. All
systems in the universe can subsequently be regarded as elements,
representing equally many biological/physical sub-systems.
(b) Biological and physical systems
When regarding systems as biological or physical this implies that these
systems show a material and/or energetic existence. According to this focus, the
concepts of wood and marble are excluded from the approach, while the specific
brain states that are associated with a human's thought about the categories
`wood' or `marble' are included. This excludes from the analyses any hierarchies
104
Chapter 4
that are based on temporal, spatial or symbolic aspects such as duration
(seconds, minutes, hours, etc.), separation (various measures for lengths,
surfaces, volumes, etc.) and numbers (1, 2, 3, etc.).
(c) Hierarchy
The hierarchy concept relates to an ordering of entities into a sequence that is
based on a relation that shows three properties (e.g. Bunge, 1969; Simon,
1973): (1) It is transitive, which means that if a has a lower hierarchical position
than b and b has a lower position than c, then a has a lower position than c. (2)
It is irreflexive, which means that a can never hold a hierarchical position below
itself (3) It is asymmetrical, which means that if a holds a lower position than b
then b cannot hold a lower position than a. The latter implies that as soon as a
group of entities shows a circular relationship, one must consider them as
having `stepped out' of the particular hierarchy. Elements showing a circular
relationship require a new way of analysis, basing hierarchical considerations on
the relationships between different groups of circularly related elements.
In system science, the importance of a circular pattern of relationships has long
been underestimated. It is only recently that an increasing body of literature has
arisen emphasising the importance of circular interaction patterns for
recognising elements and hierarchy. Such publications include discussions of
the hypercycle (Eigen & Schuster, 1979; Kauffman, 1993), emergent
organisation (Laszlo, 1996; Ponge, 2005), major evolutionary transitions
(Maynard Smith & Szathmáry, 1995a, 1995b), meta-system transitions that are
regarded as the quanta of the evolution of complexity (Turchin, 1977), relational
closure (Heylighen, 1989a, 1990), closure in different scientific contexts (Bunge,
1992; Chandler & de Vijver, 2000) and the operator hierarchy (Jagers op
Akkerhuis & van Straalen, 1999; Jagers op Akkerhuis, 2001).
The concept of hierarchy has a long history and has been applied in many
different ways and situations of which a few examples will be discussed
presently. A well-known approach to system analysis uses the three-level
hierarchy that includes the world, the system and its elements. Applying this
approach in an iterative way, the former system and element become the new
world level and system level of the next analysis. In different forms, this threelevel approach can be recognised in theoretical publications, for example, a
review of principles of hierarchy theory by Feibleman (1954), the holon approach
that was proposed by Koestler (1978), a hierarchy of system levels by Varela
(1979) and an approach based on doublets by Jaros & Cloete (1987).
The literature offers specifications of various aspects of hierarchy (e.g. reviews
by Klijn, 1995; Valentine, 2003). If a higher level in a hierarchy consists of
physically joined elements, like parts of an alarm clock or cells in a multicellular
organism, this represents a constitutive hierarchy (Mayr, 1982). If the elements
A backbone for system science
105
are not physically connected, but associated in a series of increasingly inclusive
entities, such as organisms in a population, this is considered an aggregative
hierarchy (Mayr, 1982). If elements of levels lower than the next-lower level
contribute to a certain level in a hierarchy this represents a cumulative hierarchy.
An example is the cumulative constitutive hierarchical organisation of
multicellular organisms, in which, for example, bone and blood plasma, which do
not consist of cells, together with tissues and organs form the organism. The
cumulative constitutive hierarchy in organisms has also been called a somatic
hierarchy (Eldredge, 1985). If low levels in a hierarchy represent systems that
are separated in time from the higher levels and do not function as units in the
higher levels, this is called a tree (Valentine, 2003). In addition to being
hierarchic, trees are defined as having a single root and showing a single parent
for each node. For this reason, the parent-offspring relationship (family tree) is
not strictly a tree, but more a treelike network. The pedigree of species that
forms the phylogenetic tree or `tree of life' can be considered a tree as long as
the speciation is based on a representation of the gene-pool of a species as a
single parental node. Networks or webs, then, may show nodes with
connections to a variable number of other nodes irrespective of their hierarchical
level.
(d) Mechanisms
If a pattern occurs in a biological or physical system, it always shows a
relationship with some sort of underlying process or explaining mechanism
causing it.
The most general mechanism in nature is the fact that spontaneous processes
are associated with a decrease of energy gradients and increasing
chaos/entropy. This general principle does leave room for a system to move in
the opposite direction by showing a local increase in its organisation (associated
with a decrease in entropy) as long as the related entropy decrease is
compensated for by an equal or larger increase outside the system (e.g.
Prigogine & Stengers, 1984).
Another mechanistic aspect is the relative stability of a system. From any two
systems showing an equal chance of formation, the system that shows the best
combination of internal stability and stability during interactions with other
systems will show the highest chance of functioning successfully and existing for
a long time.
Still another mechanistic aspect is the self-organisation of systems in response
to certain attractor states. Selforganisation implies that interactions between
systems autonomously create patterns. The most important aspect of selforganisation in nature is the formation of circular interaction patterns creating
physical units. Although there may be different mechanisms behind the
106
Chapter 4
emergence of atoms or cells, the occurrence of a circular interaction pattern is a
constant. This aspect will be discussed in detail in the next section.
(2) A strict basis for the hierarchical ranking of system types
For later analysis of hierarchy in the internal organisation of systems, a special
approach is used that recognises a strict division of all systems into two major
groups.
The systems of the first group are discussed in detail by Jagers op Akkerhuis &
van Straalen (1999) and Jagers op Akkerhuis (2001), who refer to these
systems as the operators, indicating clearly that these systems show a specific,
internal organisation allowing them to operate as individuals and produce effects
in their environment. The name operator was chosen, even though it was
realised that this name could cause confusion because it has applications in
other fields, such as mathematics, the telephone business and information
science.
Systems that are not operators belong to the second group, the members of
which can be regarded as interaction systems.
Because the operator theory forms an essential aspect of the present study, we
first present a short summary. According to the operator viewpoint, an operator
of type x creates a next operator of type x+ 1 by means of a first-next possible
closure. Closure refers simultaneously to the formation (the closing process) and
presence (the closed state) of a circular pattern in the interactions between the
system's elements. The adjective 'first-next possible' refers to the demand that
the closure must be the first possibility for a new type of closure in system x + 1
after the preceding closure created the operator of type x. The demand of firstnext possible closure implies that the elements showing this property can be
ranked in a strict way, creating what has been called the operator hierarchy, in
which every operator holds a unique hierarchical position (Fig. 4.1). In the
operator hierarchy, there are two transitions between system types that are
based on first-next closures: the major and the minor transitions (Jagers op
Akkerhuis, 2001). A major transition creates a completely new type of closure.
According to the operator theory, major transitions form the basis of the
superstring, the quark-gluon plasma, the hadron, the atom, the cell and the
organism with a hypercyclic neural network with interface, which is named a
memon in the operator hierarchy. A minor transition recreates a system property
that came into existence during a preceding major transition. For example, the
multiparticle property that emerged as the result of a major transition in the
hadrons occurs again as the result of a minor transition at more elevated levels
in the hierarchy, creating the molecule, the prokaryote multicellular organism
and the eukaryote multicellular organism.
A backbone for system science
107
Using a sensu stricto interpretation of the operator hierarchy, only the systems in
Fig. 4.1 showing a hypercycle with interface represent operators. This includes
the hadron, the atom, the multi-atom (e.g. molecules, metal grids, etc.), the
bacterial cell and the bacterial multicellular, the eukaryote cell and the eukaryote
multicellular, and the memon. For additional information about the operator
hierarchy, see Jagers op Akkerhuis & van Straalen (1999), Jagers op Akkerhuis
(2001) and the present author's website www.hypercycle.nl.
The operator theory may have marked effects on the analysis of organism types
in biology. As Fig. 4.2 shows, all species of organisms and the representation of
their ancestral tree (the tree of life) can be translated into a sequence of operator
types including bacterial cells, eukaryote cells, prokaryote and eukaryote
multicellulars and organisms with hypercyclic neural networks (memons).
Versions of the scheme of Fig. 4.2 that exist in the literature (e.g. Alberts et al.,
1989) generally include only the levels of prokaryote unicellulars, eukaryote
unicellulars and eukaryote multicellulars. Because the operator hierarchy offers
strong arguments to regard the transition towards multicellular eukaryotes with a
hypercyclic neural network (memons) as of similar importance as the transition
from prokaryote to eukaryote cells and from uni- to multicellular organisms, we
suggest including the level of the memon in this type of analysis.
As stated above, the operator hierarchy allows a distinction between the
operators and all other systems, regarded as interaction systems. It will be
shown next that this distinction is a useful tool for the analysis of hierarchic
organisation in systems. As an example, let us discuss the simple situation of a
system (S) that contains a water molecule (M) and a water droplet (D) (Fig. 4.3).
The operator hierarchy regards the water molecule an operator and the droplet
and system S interaction systems. The droplet represents an interaction system
because it consists of interacting water molecules that do not show first-next
possible closure, which, following the molecular stage, is autocatalysis; a
property not shown by D. The system S also represents an interaction system,
because it contains elements the interactions of which do not show first-next
closure.
108
Chapter 4
structural
autoevolution
structural
copying of
information
hypercycle multi-systems
mediating
interface
hypercycles
inter-faces
memon
eukaryote
cell
eukaryote
multicellular
CALM
sensors
-
cell
atom
multiatom
hadron
autocatalytic set
membrane
nucleus
electron
shell
superstring
plasma
confinement
major transitions
minor transitions
superstring
Figure 4.1. The ranking of system types according to the operator hierarchy Jagers op
Akkerhuis & van Straalen, 1999; Jagers op Akkerhuis, 2001). Grey boxes indicate nonoperator systems that play an important role in the operator hierarchy as intermediate
closure states. Black upward arrows represent major transitions creating a new operator
that shows a completely new type of closure. Black right-pointing arrows represent minor
transitions. Empty cells and dashes indicate stages that have not yet evolved, but
according to the logic of the hierarchy may potentially exist. Systems in the same vertical
column share a common closure type. Titles above the columns indicate closure types.
`Interface' represents an emergent boundary. `Hypercycle' represents an emergent
A backbone for system science
109
second-order interaction cycle. `multi-operator' represents an emergent recurrent
interaction between operators of the preceding type. 'Hypercycle mediating interface'
(HMI) represents an interface that mediates the interactions of the hypercycle of the
system involved with the world. `Structural copying of information' (SCI) represents the
property of systems to autonomously copy their structure and in this way reproduce their
information. `Structural auto-evolution' (SAE) represents the property of systems to
improve, while living, the neural structures that contain their information. CALM stands for
a Categorizing And Learning Module, representing a hypercyclic neural interaction
pattern.
Figure 4.2. A schematic representation of the phylogenetic tree. Organisms are ranked
according to speciation patterns, at the same time indicating when a specific lineage
passes through one of the major levels of structural organisation recognised by the
operator hierarchy
For later discussion, it is practical to distinguish two major types of interaction
systems: compound objects, in which elements by their interactions create a
material unit, and interaction groups, in which elements interact as separate
material units. The water droplet in our example can be regarded as a
compound object. Any molecule that escapes from it becomes a separate
operator: a water molecule.
Compound objects always show one or more unifying forces between one or
more types of contributing elements that create a stronger coherence between
the particles in the compound object than between the object and its
110
Chapter 4
environment. At the point where the influence of these forces comes to a halt,
the compound object has its limit that forms the basis for its distinction and
manipulation as a material unit relative to its environment. The limit makes the
compound object recognizable in space and time and may cause specific
emergent properties (Ponge, 2005). Examples of compound objects are: a drop
of water (in oil or in air, but not in water, because a water droplet in water neither
shows a recognizable outer limit nor shows specific unifying forces), a planet, a
bowl with soup, a lump of clay, a piece of dead wood, a brick, a lump of small
magnetic parts of iron clinging together on a table, a piece of cotton cloth
(unified by the molecular forces between molecules in the cotton fibres and by
physical forces keeping together the interwoven strands of fabric), a hair, etc.
Interaction groups, then, consist of particles that show specific interactions that
make them recognizable as a group, without this leading to any physical unity.
Examples are: atoms and molecules in a gas, a tornado, a heap of loose sand
particles (not kept together by roots, fungal hyphae or such like), organisms of a
population/ species in an ecosystem, bees belonging to one hive, an
autocatalytic set in a chemical solution, stars and planets in a galaxy, a football
team, etc.
Although operators, compound objects and interaction groups show distinct
types of organisation, they have all evolved from operators and may respond to
changing conditions by a change in organisation from one type to the other. For
example, a group of loosely interacting atoms in a gas (an interaction group)
may condense to form a compound object, such as a drop of rain or a
snowflake. The atoms may separate again when the system is heated. Clearly
also a change in environment may alter the status of a system. For example, a
drop of water is characterised as a compound object in air, but not in water.
It is recognised that between every pair of subsequent system types, transition
states may exist that cannot be classified as representing one of the two system
types. Transition states are also a natural phenomenon during the formation of a
higher-level system type from elements, for example, during the creation of
operators. In my opinion, the existence of transition states does not justify the
rejection of organisational system classes.
(3) Viewpoint dependence of hierarchy: the DICE approach
The operator hierarchy has been introduced as a special way of ranking a
limited subset of natural systems in a strict hierarchy. It was also shown that
systems that are not operators can be classified as different forms of interaction
systems, either compound elements or interaction groups. This provides a basis
for the analysis of the internal organisation of both operators and interaction
systems. The internal organisation of these systems can be looked at from two
perspectives: that of the elements and that of the interactions.
A backbone for system science
111
S
H2O
H2O
H2O
H2O
H2O
M
H2O
D
Figure 4.3. A system S containing a water molecule M and a drop of water D and water
molecules (H2O).
(a) Elements
A general analysis of system organisation must account for the fact that the
elements in a system are not always operators, but may also be compound
elements and interaction groups. For example, the internal organisation of a dog
(an operator of the memon type) includes interaction systems, such as organ
systems, organs and tissues, and operators, such as many specialised
eukaryote cells. Likewise, an interaction system, such as a galaxy, includes
interaction systems, such as the stars, planets, comets, dust particles, etc., and
operators, such as atoms, molecules, cells, etc. If the operators, compound
elements and interaction groups are defined, the identification of elements in a
system will not be problematic.
(b) Interactions
A general analysis must acknowledge that relationships of elements in a system
are the result of many types of interactions, each of which may lead to different
patterns of relationships among elements. To make this `many patterns problem'
tangible, interactions can be arranged according to a limited number of
`organisational dimensions'.
For example, a focus on the relationship `a causes the displacement of b' can be
regarded as one of the rankings of organisms according to a displacement
dimension. Displacement may furthermore relate to interactions involving
migration, phoresy, endochory, etc. Interactions between the same individuals
will sort differently when focusing on the relationship `a has genes that are used
by b'.
112
Chapter 4
This relationship represents one aspect of the information dimension.
Information may furthermore include interactions involving speciation, life
histories, behaviour and communication. Rankings will again change by focusing
on constructional interactions. Construction includes the way objects are
arranged in a system, the creation of objects by individuals, constructional
aspects of phenotypes and the contribution of chemicals in food to the
construction of an organism, for example, by means of vitamins, proteins or
toxins. Finally, the relationship `a eats b' is an example of an energetic
viewpoint. Other aspects of the energetic dimension may include behaviour
aiming at maximising resource dominance (e.g. Jagers op Akkerhuis &
Damgaard, 1999), the metabolism of an organism, energy flows in a food chain
and physiological effects of temperature, light, etc. As has been emphasised by
Arditi, Michalski & Hirzel (2005) the structure of a food-web may differ markedly
from a construction-chain, indicating that each of the above dimensions will
result in a different analysis of relationships and hierarchy in a system.
Together these four different dimensions will henceforth be summarised by the
acronym DICE (displacement, information, construction and energy). It is argued
here that until all the four DICE-viewpoints have been investigated, the analysis
of the organisation of any system is principally incomplete.
To summarise the above, there are important practical consequences of the
DICE approach. Firstly, it shows that each viewpoint that is used for analysing a
system results in another arrangement of relationships. Secondly, the
dimensions of the DICE approach offer an easy way to check whether the
analysis of a certain system shows major flaws.
(4) Systems that change operator type during their existence
The operator hierarchy also has consequences for the analysis of system types
that occur during development. For example during conception, the unicellular
organisms of the sperm cell and egg cell fuse to form a zygote, which also
represents a unicellular organism. Development can now proceed along different
lines. In species with determinate cleavage of the zygote, the blastomere cells
depend immediately on each other for their survival and are never separate
individuals. In other taxa, for example many mammals, the blastomere cells
specialise much more slowly and have the potential to develop into individual
multicellular organisms when separated. Accordingly, the mammalian zygote
represents a colony of structurally linked cells. In a later phase, the cells become
obligatorily interdependent and the colony changes into a multicellular organism.
In animals, the development of the embryo passes through a stage where a
hypercyclic neural network forms. From that moment on, the embryo becomes a
memon. Other interesting examples are slime moulds. Individual cells of these
organisms may live as separate individuals and even multiply asexually at this
A backbone for system science
113
stage. On certain occasions, the cells aggregate and form slug-like units, which
show all properties of a multicellular stage, with chemical bonds between their
cells and a mutual dependence with respect to a common metabolism and
reproduction.
Above, I discussed transitions between complexity levels in operators. However,
transitions between complexity levels also occur in interaction systems. As an
indicator for the complexity level, we proposed the use of the highest level
operator in an interaction system. According to this viewpoint, a planet starts its
life as a chemosystem and changes towards an ecosystem when the first cell
emerges. Any subsequent level organisation of a planet can be named after
newly emerging organisms, which, for example, may lead to the recognition of
an ecosystem at the unicellular eukaryote level, at the multicellular eukaryote
level (e.g. plants and fungi) or at the memon level (includes most animals).
Analysis of hierarchy in biological and physical systems
The method proposed in the present paper includes the following four steps: (1)
identification of the developmental stage of the universe, (2) identification of the
type and scale of the system, (3) specification of mediating influences affecting
the formation of the system and, (4) analysis of the internal organisation based
on the four dimensions of the DICE-approach. These steps are explained in the
sections below.
(1) Identification of the developmental stage of the universe
As a first step, the present method assesses which types of systems must
potentially be included in the analysis. The presence of system types is
considered to depend on the succession stage of the universe. According to the
operator hierarchy, succession stages of the universe can be defined on the
basis of the highest complexity operator that is present. Based on this viewpoint,
the universe has passed through a number of abiotic stages that are associated
with the emergence of, for example, hadrons, atoms and molecules, and biotic
stages that are associated with the emergence of prokaryote and eukaryote
cells, multicellular organisms and memons. Because operators and interaction
systems show a mutual dependency with respect to their formation, the
evolutionary sequence of operators shows a correlation with the presence of
interaction systems in the universe; this relationship is illustrated in Fig. 4.4.
Although in Fig. 4.4 the presence of the highest-complexity operator determines
the succession stage, the universe may show large areas in which evolution lags
behind. Due to this heterogeneity, the universe at the stage of unicellular
eukaryotes (white shading in Fig. 4.4) may contain large parts in which cellular
life has not yet emerged and where the analysis of system organisation does not
have to take into account the activities of organisms. Furthermore, the scheme
114
Chapter 4
shown in Fig. 4.4 holds open the possibility that operators of early stages either
become parts of higher-level operators or may remain present as individual
entities through later succession stages of the universe.
(2) Identification of a system of interest
This step is used to identify the system of interest. This implies that the scale of
the system is specified in relation to its characterisation as operator or
interaction system. For example, if the selected system is a galaxy, it consists of
celestial bodies (and dark matter) kept together by gravitational interactions.
According to the present analysis, a galaxy represents an interaction group
depending on compound elements (the stars and planets). The scale for
analysing its organisation depends on judgments about the limit to the
gravitational influence it has in space. The situation becomes quite different if
the system chosen is a plant. This represents an operator of the eukaryote
multicellular type. The scale for analysing its organisation is that of the individual
and is limited by intercellular connections.
(3) Mediating influences
The third step involves a further characterisation of the system by analysing
whether the system has obtained its specific form under the influence of a
higher-level operator or because of specific interactions within a larger
interaction system. The advantage of this step lies in the fact that it makes
explicit that any analysis of the organisation of a system requires reference to
the surrounding environment and/or higher level that mediated its construction.
On the one hand, systems exist that typically owe their form to specific
interactions within interaction systems. For example, carbon atoms form during
nuclear interactions in stars, crystals form under specific environmental
conditions on planets and the canopy of a forest affects the selection of new
seedlings. On the other hand, systems may experience mediating activities of
higher-level operators. For example, a fossil and a DNA molecule could never
have obtained their present form without the mediating influence of an organism.
(4) Internal organisation
The fourth and last step is an investigation into the elements composing the
selected system and how they are related. The elements may be operators,
compound objects or interaction groups. One can analyse the internal
organisation of these elements in an iterative way, focusing on systems
elements, the elements therein, etc. When it comes to finding explanations for
the functioning of a system, it is often sufficient to go down one or two levels. As
a consequence of the DICE approach (see section 11.3) there is no one best
hierarchical ranking of interactions in the system, because the ranking of the
elements may vary with the viewpoint that is adopted for acknowledging
hierarchy.
A backbone for system science
115
Care should be taken when analysing the organisation of interaction systems.
For example, the ranking from the organism, to the population, to the community
and the ecosystem, which can be observed frequently in the literature, should
not make the reader think that organisms are first parts of populations, which
then are parts of communities, which finally are parts of ecosystems. Instead,
this hierarchy only involves abstract subsets, and not structural elements. Fig.
4.5 illustrates that the organism interacts as an individual with other svstems in
the ecosystem without creating new structural elements.
On the basis of this reasoning, a population, a community, a food chain and
other groups of individuals must be considered abstract subsets, each being
based on a specific selection of interactive properties of the individuals.
Nevertheless, even though the interactions of the individuals in such abstract
groups may not lead to physical units, they do cause very real dynamics. This
can easily be demonstrated by the example of mating and genetic recombination
in populations of species that reproduce sexually. Due to sexual reproduction,
the offspring obtain different gene combinations. On the one hand, this helps to
maintain good gene combinations in part of the offspring, the part carrying the
most deleterious mutations experiencing a survival disadvantage. On the other
hand, the recombination of genes during sexual reproduction brings about new
gene combinations that allow a flexible adaptation of a species to changing
fitness landscapes.
116
Chapter 4
Subsequent stages of the universe ->
Increasingly complex interaction systems ->
possible system types in a universe
that contains eukaryotic cells as
highest complexity level operators
interaction systems
I(S-mA)
I(S-M)
I(S-mC)
I(H-M)
I(S-eC)
I(H-mC)
I(A-M)
I(S-C)
I(H-eC)
I(A-mC)
I(mA-M)
I(H-C)
I(A-eC)
I(mA-mC)
I(C-M)
I(S-A)
I(H-mA)
I(A-C)
I(mA-eC)
I(C-mC)
I(eC-M)
I(S-H)
I(H-A)
I(A-mA)
I(mA-C)
I(C-eC)
I(eC-mC)
I(mC-M)
I(S
I(H)
I(A)
I(mA)
I(C)
I(eC)
I(mC
I(M)
S
H
A
mA
C
eC
mC
M
S(H)
H(A)
A(mA)
mA(C)
C(eC)
eC(mC)
mC(M)
S(A)
H(mA)
A(C)
mA(eC)
C(eC)
eC(M)
S(mA)
H(C)
A(eC)
mA(mC)
C(M)
S(C)
H(eC)
A(mC)
mA(M)
S(eC)
H(mC)
A(M)
S(mC)
H(M)
‘elementary’ particles
(the operator hierarchy)
S(M)
Increasingly complex operators ->
Figure 4.4. System types associated with different succession stages of the universe.
Each successive step to the right adds a box both to the operators (lower panel) and to
the interaction systems (upper panel), the boxes being pushed outward at each step.
From left to right, subsequent columns indicate all system types, both operators (bottom)
and interaction systems (top), which potentially exist during the following succession
stages of the universe. Abbreviations: S = superstring stage, H = hadron stage, A = atom
stage, mA= multi-atom stage, C = prokaryote cell stage, eC = eukaryote cell stage, mC =
multicellular organism stage (including in this case both prokaryote and eukaryote
multicellulars) and M =memon stage. The coding I(X-Y) indicates all possible interaction
systems containing type X as the highest-level operator in a succession stage of the
universe that contains Y as its highest-level operator. For example, the coding I(eC-M),
covers all possible interaction systems that show eukaryote cells as the highest level
elements in a universe in which memons exists. For the operators the coding X(Y) is used
to indicate all possible operators of type X in a stage of the universe that contains Y as its
highest-level operator.
A backbone for system science
117
Applications of the new method
The application of conflicting hierarchy rules and the inclusion of ill-defined
systems flaw many existing analyses of hierarchy in system organisation. To
solve this problem, an alternative method is proposed in the present article.
Below, possible contributions of the new method to the field of system analysis
are discussed.
(1) Distinguishing between evolutionary sequence and construction
sequence
Many hierarchic approaches to system organisation use an organism concept
that covers all types of organism and link this to an internal organisation of
organisms that includes organs, tissues and cells. This viewpoint disregards
fundamental differences in complexity among organisms, thus failing to give an
accurate analysis of their internal organisation. To avoid such problems, the
present approach uses the operator hierarchy to determine (and rank) the
complexity level of the organism before analysing internal organisation. This
approach also solves the problem that ranking of systems according to an
evolutionary sequence does not always correspond with ranking according to
complexity. For example, before the emergence of the cell, evolution had neither
the means nor a context for developing complex organelles, such as the
endoplasmic reticulum. Therefore, cells evolved first, followed by organelles.
The evolution of internal complexity has been discussed by Turchin (1977) who
refers to it as the `law of the branching growth of the penultimate level'. This law
states that only after the formation of a control system C, controlling a number of
subsystems Si, will these S; tend to multiply and differentiate. Examples of
elements that have evolved in organisms as indicated by Turchin's law are
organelles in cells and tissues, organs and specialised cells in multicellulars.
(2) Classifying and analysing systems in relation to their creation under
the influence of higher-level operators
Analyses of the organisation of systems generally do not include mediating
effects of higher-level systems and the environment. The present approach
deals specifically with this aspect, increasing insight into the organisation of
systems.
(3) Adapting the scale of the systems of interest in relation to specific
interactions
Most analyses in the literature do not explicitly consider the scale of the systems
involved. Populations, communities and pastoral systems are all mentioned
without any specification of what sets the limits to these selections.
118
Chapter 4
ecosystem
local
community 1
local
community 2
sub-population A
subpopulation B1
subpopulation B2
individuals of predator (species A)
individuals of prey (species B)
Figure 4.5. The partitioning of individuals into subsets of an ecosystem. Black arrows
indicate interactions between individuals. Arrows marked with a cross indicate sets that
do not interact as entities, such as populations and communities. The figure shows that it
is always the individual (or a physically connected colony of individuals that acts as one
individual) that interacts in the ecosystem, regardless of the subset it is assigned to.
(4) Specification of the type of hierarchy used in different hierarchical
steps
In many existing hierarchies, it is not clear which properties determine the
hierarchical ranking of any next level. Frequently, the ranking from sub-atomic
particles to the universe gives the impression of an internal hierarchy, in which
the top-down relationship `is a part of' seems to fit most of the hierarchical steps.
Galaxies are parts of the universe, solar systems parts of galaxies, planets parts
of solar systems, ecosystems parts of planets, communities parts of
ecosystems, populations parts of communities, organisms parts of populations,
cells parts of organisms, organelles parts of cells, molecules parts of organelles,
atoms parts of molecules, etc. Yet, apart from additional minor inconsistencies,
the latter hierarchy is inconsistent because it is constructed from three distinct
parts, the ranking of which is based on very different principles.
The first part involves the internal organisation of the universe down to the level
of planets. This range is based on the general notion that smaller systems are
parts of larger systems. Existing hierarchies of this type generally do not include
A backbone for system science
119
all types of celestial bodies, such as black holes, neutron stars, brown dwarfs,
comets, etc. Moreover, according to this reasoning, there is no consensus on
how to distinguish between a lifeless planet and a planet inhabited by
organisms.
The second part involves subsets of the ecosystem ranked from the organism,
via the population, to the community and the ecosystem. As has been discussed
above, organisms are not first parts of populations, which then are parts of
communities, which finally are parts of ecosystems. Instead, as was illustrated
by Fig. 4.5, the organisms remain at any moment directly integrated into the
ecosystem. It was also discussed above that the ranking of individuals in
ecosystems is sensitive to point of view, as illustrated by the DICE approach, the
application of which may result in a food web, a structural dependence web, an
informational web, and so on.
The third part involves the internal organisation of elements in the organism.
Considering the organism as just another operator, this part can also be
generalised to relate to the internal organisation of operators. The present
method covers this aspect in far more detail than most other methods. Using the
proposed rationale, the analysis can be based on the recognition of internal
elements, such as operators, compound elements and interaction groups.
Accounting for the DICE discussion, several internal hierarchies can be
recognised.
Finally, I would like to refer to the existing controversy about the usefulness of
hierarchy in system science (e.g. Webster, 1979). On the one hand Guttman
(1976), advocates that the use of levels of biological organisation
`if stated in any but the sloppiest and most general terms is a useless and
misleading concept'. On the other hand, Weiss (1969) remarks that `the principle
of hierarchic order in living nature reveals itself as a demonstrable descriptive
fact' and Von Bertalanffy (1968) that ...'hierarchical structure ... is characteristic
of reality as a whole and of fundamental importance especially in biology,
psychology, and sociology'.
The present study brings together these opposing viewpoints. On the one hand,
it shows that it is indeed difficult to use hierarchy as a scientific concept. This will
continue as long as approaches focus on ill-defined hierarchies that include
various system types and ill-defined hierarchy rules. On the other hand, the
present study shows that hierarchy can be studied with success and in detail by
using the operator hierarchy as the basis at the same time paying close attention
to the multi-dimensional nature of hierarchy in biological/physical systems.
120
Chapter 4
Conclusions
(1) The literature shows a controversy about the usefulness of hierarchy in
analysing the organisation of biological/physical systems. On the one hand,
it is postulated that hierarchy is the most general organising principle in
nature. On the other hand, the identification of hierarchy in natural systems
seems to be hampered by a sloppy use of concepts, giving reason to claims
that hierarchy is of limited use. Especially linear hierarchies can be shown
to suffer from minor and major flaws.
(2) Solving the above problems requires a strict yet flexible way for analysing
system organisation. With the operator hierarchy as a basis, we propose a
method that includes the following four steps: (a) identification of the
developmental stage of the universe, (b) identification of a system of
interest, (c) analysis of mediating influences on the selected system, (d)
analysis of internal organisation.
(3) The present method of identifying contributes in three ways to the analysis
of system organisation: (a) it offers a strict ranking of the operators, (b) it
offers ways to identify compound elements and interaction groups, and (c) it
acknowledges that the analysis of hierarchy in interactive relationships must
focus on different analytical dimensions, notably displacement, information,
construction and energy.
Acknowledgements
The author would like to thank two anonymous referees, as well as Hans-Peter
Koelewijn and Jean-François Ponge for detailed comments on the manuscript,
Chris KIok and Jan Klijn for constructive discussion and Claire Hengeveld for
improving the structure and clarity of the text and correction of the English.
Chapter
5
Bringing the definition of life to closure
Towards a hierarchical definition of life, the
organism and death
G.A.J.M. Jagers op Akkerhuis
Foundations of Science, Volume 15, Number 3: 245-262, 2010, in press
G.A.J.M. Jagers op Akkerhuis
Explaining the origin of life is not enough for a definition of life.
Foundations of Science (forthcoming), 2010, in press
“Life is a succession of lessons which we must live to be understood.”
(Ralph Waldo Emerson)
122
Chapter 5
Abstract
Despite hundreds of definitions, no consensus exists on a definition of life or on
the closely related and problematic definitions of the organism and death. These
problems retard practical and theoretical development in, for example,
exobiology, artificial life, biology and evolution. This paper suggests improving
this situation by basing definitions on a theory of a generalised particle
hierarchy. This theory uses the common denominator of the ‘operator’ for a
unified ranking of both particles and organisms, from elementary particles to
animals with brains. Accordingly, this ranking is called ‘the operator hierarchy’.
This hierarchy allows life to be defined as: matter with the configuration of an
operator, and that possesses a complexity equal to, or even higher than the
cellular operator. Living is then synonymous with the dynamics of such operators
and the word organism refers to a select group of operators that fit the definition
of life. The minimum condition defining an organism is its existence as an
operator, construction thus being more essential than metabolism, growth or
reproduction. In the operator hierarchy, every organism is associated with a
specific closure, for example, the nucleus in eukaryotes. This allows death to be
defined as: the state in which an organism has lost its closure following
irreversible deterioration of its organisation. The generality of the operator
hierarchy also offers a context to discuss ‘life as we do not know it’. The paper
ends with testing the definition’s practical value with a range of examples.
Keywords: Artificial life, biology, evolution, exobiology, natural sciences, particle
hierarchy, philosophy, Big History
On life and death
123
Introduction
In a chronological overview of developments, Popa (2003) presents about 100
definitions of life, meanwhile demonstrating that no consensus exists. Many
classical definitions include long lists of properties, such as program,
improvisation, compartmentalisation, energy, regeneration, adaptability and
seclusion (Koshand Jr., 2002) or adaptation, homeostasis, organisation, growth,
behaviour and reproduction (Wikipedia: Life). Most properties in such lists are
facultative; it is still possible to consider an organism a form of life when it does
not grow, reproduce, show behaviour, etc. The inclusion of facultative aspects is
a source of lasting difficulty in reaching consensus on a definition of life.
Because of the seeming hopelessness of the situation, certain scientists have
adopted a pragmatic/pessimistic viewpoint. Emmeche (1997) christened this
viewpoint the “standard view on the definition of life’. He suggests that life
cannot be defined, that its definition is not important for biology, that only living
processes may be defined and that life is so complex that it cannot be reduced
to physics. Others have warned that a comprehensive definition of life is too
general and of little scientific use (e.g. van der Steen 1997).
In their search for a definition, other scientists have focused on properties that
are absolutely necessary to consider an entity life. In this context Maturana &
Varela (1980, p. 78) have proposed the concept of autopoiesis (which means
‘self making’). They use the following definition: “An autopoietic machine is a
machine organised (defined as a unity) as a network of processes of production
(transformation and destruction) of components which: (i) through their
interactions and transformations continuously regenerate and realise the
network of processes (relations) that produced them; and (ii) constitute it (the
machine) as a concrete unity in space in which they (the components) exist by
specifying the topological domain of its realisation as such a network.” Special
about the autopoietic process is, that it is “closed in the sense that it is entirely
specified by itself (Varela 1979 p. 25)”.
The concept of autopoiesis has increasingly become a source of inspiration for
discussions in the artificial life community about how to define life (Bullock et al.,
2008). Reducing the number of obligatory traits defining life to just one,
autopoiesis is a rather abstract concept. People have sought, therefore, to
describe some of the processes that underlie autopoiesis more specifically. An
example of such a description is a triad of properties defining cellular life:
container (cell membrane), metabolism (autocatalysis) and genetic program
(e.g. Bedau, 2007).
124
Chapter 5
These descriptions, however, have not resulted in a consensus definition of life.
This has led Cleland & Chyba (2002, 2007) to suggest that a broader context, a
‘theory of life’, is required. In line with a broader framework, life may be regarded
as a special realisation of the evolution of material complexity. According to
Munson and York (2003), considering life in a general evolutionary context
requires arranging “all of the phenomena of nature in a more or less linear,
continuous sequence of classes and then to describe events occurring in the
class of more complex phenomena in terms of events in the classes of less
complex phenomena.. “. An important property of such a hierarchy would be that
“…an increase in complexity is coupled with the emergence of new
characteristics … suggesting that the hierarchical arrangement of nature and the
sciences is correlated with the temporal order of evolution”. Similar views for
integrating material complexity and the evolution of life can be found, for
example, in the work of Teilhard de Chardin (1966, 1969), von Bertalanffy
(1968), Pagels (1985), Maynard Smith & Szathmáry (1995, 1999, 2002) and
Kurzweil (1999).
In contribution to these discussions, the present author has published an
evolution hierarchy for all ‘particles’. The latter hierarchy uses the generic word
‘operator’ to address both physical (e.g. quark, atom, and molecule) and
biological particles (e.g. prokaryote cell, eukaryote cell, and multicellular). The
word operator emphasizes the autonomous activity of the entities involved,
which ‘operate’ in a given environment without losing their individual
organisation. The hierarchical ranking of all operators is called the ‘operator
hierarchy’ (see Figure 5.1) (Jagers op Akkerhuis and van Straalen 1999, Jagers
op Akkerhuis, 2001, Jagers op Akkerhuis, 2008 and the author’s website
www.hypercycle.nl). Because the operator hierarchy is important for the
definition of life proposed below, the outlines of this theory are summarised in
the following lines
The operator hierarchy ranks operators according to the occurrence of a circular
pattern, such as that which connects the beginning and end of a process or
structure. Circularity causes a closed organisational state, also referred to as
‘closure’ (for discussions of closure see, for example, Heylighen 1990, Chandler
and van de Vijver, 2000). Because closure causes a discrete ‘quantum’ of
organisation (e.g. Turchin 1977, 1995 and Heylighen 1991), the operator
becomes an ‘individual entity’, a ‘whole’ or a ‘particle’, while still retaining its
construction of smaller elements. Closure thus defines the operator‘s complexity
level and sequential closures imply a higher complexity level. An operator‘s
closure is the cause of its existence and typical for its complexity. This implies
that complexity is not measured in terms of the number of genes, functional
traits or organs of an organism, but in a very abstract way, in terms of the
number of closures. Upon losing its closure, the organisation of the operator falls
back to that of the preceding operator. The actual shape of a closure can differ.
On life and death
125
Biological examples of closure are the cell membrane and the circle of catalytic
reactions allowing the cell to maintain its chemical machinery. It is essential for a
strict ranking that a lower-level and a higher-level operator always differ by
exactly one closure level. The single closure (eukaryotic cell) or parallel pair of
closures (autocatalysis plus membrane of the cell) that define the next level are
referred to as ‘first-next possible closure(s)’. A consequent use of first-next
possible closures allows physical and biological operators to be ranked
according to the ‘operator hierarchy’ (Fig. 5.1). The operator hierarchy includes
quarks, hadrons, atoms, molecules, prokaryotic cells, eukaryotic cells,
multicellulars (e.g. plants, fungi) and ‘animals’, the latter representing an
example of the operators that possess a neural network with interface and that
are called ‘memons’ in the operator hierarchy.
Due to its focus on closure, the operator hierarchy represents an idealisation
because it excludes potential transition states in between two closures. For
example, several hundreds of metal atoms may be required before a functional
Fermi sea transforms a collection of single atoms into a metal grid. Also, the
emergence of multicellularity (discussed in detail below in the ‘Levels of life’
chapter) may require a colonial, multicellular state in between the single cell and
the multicellular operator. The above shows that transition states form natural
intermediate phases in the emergence of closures. The operator hierarchy does
not include these transition states, however, because its hierarchical ranking is
exclusively based on entities that already show first-next possible closure.
The main reason for writing this paper, and adding yet another definition of life to
the listings, is that the operator hierarchy offers several advantages in solving
definition problems. First, the definitions of the operators are generally
applicable because they focus on the essences of organisation. For example,
demanding autocatalysis leaves open which specific catalysts will perform the
process. Second, the use of first-next possible closures ensures a critical
filtering of only obligatory properties from property lists. Finally, the use of the
operator hierarchy makes it easy to develop a hierarchy-based definition of life.
In other words, the operator hierarchy offers a novel path for structuring and
simplifying discussions about which entities are life.
126
Chapter 5
memon-based
operators
EUKARYOTE
MULTICELLULAR
MULTICELLULAR
cell-based
operators
EUKARYOTE
CELL
CELL
ATOMS
MOLECULES
HADRONS
atom-based
operators
MEMIC
ORGANISMS
CELLULAR
ORGANISMS
(HARDWIRED)
MEMON
hadrons
OPERATORS
REPRESENTING THE ‘DEAD’
STATE OF MATTER
OPERATORS WITH A STATE OF
MATTER REPRESENTING LIFE
THE OPERATOR HIERARCHY
Figure 5.1. Using the operator hierarchy to define life and organisms. Arrows indicate
how closures create operators (more information can be found in Jagers op Akkerhuis
2008, and the author’s website www.hypercycle.nl).
The following paragraphs discuss different aspects of existing definitions of life
and examine new ways to define the organism, living and death. At the end, a
test of the practical value of the present definitions for the solving of a range of
classical problems, such as a virus, a flame, a car, a mule and a mitochondrion,
will be presented.
On life and death
127
Defining life and the organism
Before discussing the use of the operator hierarchy for defining life, living and
the organism, it is important to note that when talking about definitions, care
should be taken that “a definition is a series of superimposed language filters
and only the definiendum (the term to be defined) can penetrate it” (Oliver and
Perry, 2006). Problems may arise when the words used for the definiendum and
for the filter have a broad meaning or have different meanings in different
contexts. It is thus useful to elaborate on the current context for ‘life’ before
continuing.
‘Life’ has different meanings in different contexts. For example, people refer to
the period between birth and death as their life (this is the best day of my life)
even though lifetime would be more correct. In addition, the experience of ‘being
alive’, or ‘living’, also carries the label of life (to have a good life). Other uses of
life holistically refer to the importance of selective interactions in ecosystems that
over generations lead to better-adapted life forms (the evolution of life). RuizMirazo et al. (2004) have proposed a definition of the latter type. They state that
life is “a complex collective network made out of self-reproducing autonomous
agents whose basic organisation is instructed by material records generated
through the evolutionary-historical process of that collective network”. In
philosophy, life is sometimes considered a graded concept for being because all
what is, is alive in the measure wherein it is (Jeuken 1975). Due to the
contextual dependence of these and other interpretations, it is improbable that a
general definition of life can be constructed. Van der Steen (1997) indicates that
even if such an overly general definition existed, it would probably be difficult to
apply it to specific situations.
To avoid problems with generality and multiple interpretations of concepts, the
present study adopts a limited viewpoint, presuming a one-to-one relationship
between a definition of life and a specific material complexity. In this context, life
is an abstract group property shared by certain configurations of matter.
The operator hierarchy offers a context for a general matter-based definition of
life. Focusing on all operators showing a complexity that exceeds a certain
minimum level, the hierarchy suggests a definition of life sensu lato as: matter
with the configuration of an operator, and that possesses a complexity equal to
or even higher than the cellular operator. Only the prokaryote cell, the eukaryote
cell, the prokaryote and eukaryote multicellular, the hardwired memon and the
potential higher-level operators fit this definition (Fig. 5.1). In addition to this
general definition, various specific definitions are possible by focusing on
operators that lay between a lower and an upper closure level. An example of a
128
Chapter 5
specific definition is one describing cellular life (e.g. algae, plants and fungi) as:
matter showing the configuration of an operator, and that possesses a minimum
complexity of the cellular operator and the maximum complexity of a multicellular
operator. The latter includes only the cell, the eukaryotic cell, the prokaryotic and
the eukaryotic multicellular. It is possible to choose any of these approaches for
defining living as: the dynamics of an operator that satisfies the definition of life.
The above approach results in a strictly individual based definition of life as a
group property of certain operators. This definition has the advantage, that it
offers a solid basis for defining the creation of offspring. Subsequently, the
evolution of life can be dealt with as an emergent process occurring in any
system with interactions between individual living entities that lead to differential
survival of variable offspring, produced either without or with recombination of
parental information.
The organism is the key ontological unit of biology (Etxeberria, 2004,
Korzeniewski, 2004) and is also referred to as a ‘living individual’. Understanding
the latter requires insight into what is ‘living’, and what is an ‘individual’. By
defining ‘living’ as the dynamics of those operators that satisfy the definition of
life, the operator hierarchy uses operators instead of individuals because
operators define a being or an individual more strictly than the Latin concept of
individuum. The word individuum stands for an “indivisible physical unit
representing a single entity”. This definition leaves a great deal of room for
choice of the elements that form the physical unit and for the rules that
determine indivisibility. These indeterminacies may be the reason for the
discussion about whether certain life forms are organisms. Townsend et al.
(2008) use the phrase ‘unitary organism’ to indicate the individual organism.
However, certain jellyfish, for example, the Portuguese Man O’ War (Physalia
physalis), look like individuals, but consist of differentiated individuals, each with
its proper neural network (e.g. Tinbergen 1946). In the operator hierarchy, the
latter jellyfish are colonies, not organisms, because each contributing individual
has its proper neural network as its highest emergent property, and the colony
still lacks a recurrent interaction of the neural interfaces of the individuals.
The operator hierarchy now suggests a way to create congruency between the
definition of life and the definition of the organism by accepting as organisms
only entities that fit the operator-based definition of life. For example, using the
general definition of life, only the cells, the eukaryotic cells, the prokaryotic and
eukaryotic multicellulars and the memons are organisms.
On life and death
129
Levels of life
a. The cell. The most important properties of the cell are the autocatalytic set of
enzymes and the membrane. The autocatalytic set shows reproduction as a set.
Every molecule in the set catalyzes a reaction that produces some other
molecule in the set until any last reaction product closes the cycle. In different
ways, reproduction as a set is part of various theories about the origin of life
(e.g. Rosen 1958, 1973, 1991, Eigen 1971, Gánti 1971, Eigen and Schuster
1979, Kauffman 1986, 1993, Bro 1997, Kunin, 2000, Hazen, 2001, Martin and
Russell, 2003, Hengeveld and Fedonkin, 2007).
Autocatalysis demands that a cell can potentially autonomously sustain its
catalytic closure. Accordingly, if a cell allocates a part of its autocatalytic closure
to another cell, the cell is no longer an operator. An example of the latter is the
mitochondrion. It is generally accepted that mitochondria started the interaction
with their host cells as autonomous endosymbiontic α-proteobacteria. Over
many generations, these bacteria transferred more than 90 percent of their
catalytic control to their host (Allen 1993, Berg and Kurland, 2000, Searcy, 2003,
Capps et al., 2003, Lane, 2005). The loss of the potential of autocatalysis
implies that mitochondria have become a special kind of organelle.
In addition to autocatalysis, the operator hierarchy demands an interface
because a set of autocatalytic enzymes only gains the physical individuality that
allows its maintenance when it functions in a limited space, the limits being part
of the system. The integration of autocatalysis and the membrane is part of
various important theories, for example, the theories of autopoiesis (Varela
1979) and of interactors (Hull 1981).
b. The eukaryote cell. A single cell has two dimensions for creating a next
closure. One is to create cooperation between cells, which leads to
multicellularity. The other is to create an additional closure mediating the
hypercyclic functioning of the cell in the form of the nucleus. Interestingly, it is
quite likely that the most important complexity boundary in cell biology, that
between prokaryotic and eukaryotic cells, thanks its existence to the energy
boost and genetic enrichment offered by endosymbionts. With respect to the
emergence of eukaryotic cells, theories roughly divide along two major lines
depending on whether the nucleus or the endosymbionts emerged first. In
addition to other aspects, support for the nucleus-first hypothesis comes from
allegedly primitive eukaryotes that show a nucleus without harboring
endosymbionts. Genetic analyses (Rivera 1998) and observations of
endosymbiont traces (Clark 1995), however, suggest that the ‘primitive
eukaryotes’ are recent developments that lost their endosymbionts in a process
of evolutionary specialisation. The endosymbiont hypothesis advocates that a
130
Chapter 5
merger between a methanogenic bacterium that was member of the archaea
and an α-proteobacterial endosymbiont created the eukaryotic cell (Martin and
Russell, 2003). Subsequent transmission of genes for membrane creation from
the endosymbiont to the host allowed it to produce membranes that formed the
basis for the engulfment of the nucleus. Whatever the actual path taken by
evolution, the operator hierarchy focuses on the occurrence of closure involving
both structural and functional aspects of the host cell, resulting in an internal
interface for the autocatalytic set and the mediation of its functioning. Even
though endosymbionts may become obligatorily integrated in the functioning of
their host cell by the transfer of part of their genetic regulation to the host cell,
they do not mediate the functioning of the autocatalytic set of the host nor form
an interface for its functioning. For this reason the operator hierarchy does not
regard endosymbiosis, but the nucleus as the relevant closure that defines the
limit between prokaryotes and eukaryotes.
c. The multicellular. When does a group of cells become a multicellular operator
and, according to the above definition, an organism? In the operator hierarchy,
multicellularity involves a structural and a functional component represented by
structural attachment of cells and an obligatory recurrent pattern of functional
interactions between them. As such, it is possible to define a multicellular
operator (a multicellular organism sensu stricto) as: a construction of mutually
adhering cells showing obligatorily recurrent interactions based on the same
interaction type, that has the potential of maintaining it’s functioning as a unit
and that does not show memic structure.
Multicellularity has developed independently in many branches of the
phylogenetic tree (reviews by, for example, Bonner 1998, Kaiser, 2001,
Grosberg and Strathmann, 2007) presumably because it is associated with a
range of evolutionary advantages. Multicellularity increases mobility and access
to resources, and reduces predation, and finally yet importantly, the cells in
genetically uniform multicellulars share the same genes and do not have to
compete with each other for reproduction. Willensdorfer (2008) indicates that the
alleviation of reproductive competition allows for a division of labour because
“cells can specialize on non-reproductive (somatic) tasks and peacefully die
since their genes are passed on by genetically identical reproductive cells which
benefited from the somatic functions”.
In some cases a multicellular organism results from the aggregation of
individually dwelling unicellulars (for example, true slime moulds, Ciliates and
Myxobacteria). More generally, a multicellular organism develops when daughter
cells cohere after cell division. A simple, temporary form of multicellular life is
present in slime moulds. Here, genetically-different, individually-dwelling cells
aggregate and bind using membrane proteins to form a colonial state in which
the cells intercellularly communicate by diffusion. At a certain moment,
On life and death
131
obligatory interactions between cells lead to the formation of irreversible cell
differentiation producing a reproductive structure. During this state, the slime
mould cells are temporarily a multicellular organism.
With the evolutionary development of plasma connections, advanced
multicellular life became possible. Plasma connections allow efficient and rapid
intercellular communication, involving electrical signals, chemical signals and
nutrient transport (Mackie et al. 1984, Peracchia and Benos, 2000, Nicholson,
2003, Panchin, 2005). Plasma connections have evolved in several lineages of
multicellulars. Plasma connections between animal cells depend on gap
junctions, between plant cells on plasmodesmata, in blue-green algae on
microdesmata, and in certain fungi or in developing insect eggs on incomplete
cell walls. The evolution of gap junctions some 700 million years ago coincided
with an explosion of multicellular life forms.
Multicellular organisms may go through life stages that are not multicellular. For
example, sexual reproduction involves single-celled egg and semen.
Furthermore, during the two-, four- and early eight-cell stages most vertebrate
embryos have loosely attached cells without obligatory dependency.
Accordingly, they represent a colony. When separated from the colony, the cells
show a normal development. Early separation of embryonic cells is the reason
why identical twins exist. Embryo cells in the early stages can even mix with
another embryo‘s cells of the same age and develop into a normally functioning
organism, called a chimera, in which some organs and tissues belong to a
different genotype than others. A definition of life should, therefore, respect that
an organism‘s cells may differ in genotype. From the late eight-cell stage, the
development of gap-junctions marks the emergence of regulation as a unit,
which makes the cellular colony a multicellular.
The realisation of a multicellular‘s potential for maintenance depends on
prevailing conditions. For example, a tree twig that is stuck in the ground may
become a tree again if the weather is not too warm, too cold, or too dry, etc. and
if the twig has the genetic potential for regeneration and is large enough, in good
condition, etc.. Whether the twig is an organism depends on its potential to show
all dynamics required for being a multicellular operator. This potential is in
principle gene-based, but it depends on the condition of the phenotype and the
environment for its realisation.
Sometimes two multicellular organisms show symbiosis, such as plants living in
close association with mycorrhiza fungi in their roots. As the fungus and the
plant already are multicellular on forehand, a plant with mycorrhiza represents
an interaction between two multicellular organisms.
132
Chapter 5
d. The memon. Attempts to define life frequently focus on the typical properties
of the first cell. The underlying assumption may be that all organisms consist of
cells and that, for this reason, the definition of the living properties of cells will
automatically cover other, more complex organisations. According to the
operator hierarchy, this reasoning is incomplete because, with respect to
artificial intelligence, it unsatisfactorily excludes technical life a priori. The reason
is that the fundamental construction of the brain is not principally different when
built from cellular neurons, technical neurons (small hardware acting as a
neuron) or programmed neurons (virtual devices modelled to act as neurons).
Even though all organisms on earth currently consist of cells or show neural
networks that consist of cells, the fact that technical memons may, one day,
have a brain structure similar to cellular memons implies that a general definition
of life must consider the possibility of technical memons.
Memons show a neuron network and a sensory interface. The basic neuronunits have been named categorizing and learning modules or CALMs and allow
for a recurrent network of CALMs (Murre, Phaf and Wolters 1992, Happel 1997).
The interface includes sensors that allow the memon to perceive its body and
environment, and effectors that allow it to move the cellular vehicle it resides in.
The interface and vehicle co-evolved during the evolution of neural networks. In
principle, it is possible to construct a functional memon from any kind of
technical hardware that provides the required neural architecture. This is the
reason that the study of neural networks in biology shows a fundamental overlap
with research on technical artificial intelligence. The recognition that memons
show a recurrent network of CALMs surrounded by an interface allows Siamese
twins with separate brains to be classified as two memons sharing the same
vehicle and showing in this vehicle a partial overlap of their interfaces.
No life, no reproduction
According to some authors (e.g. the Von Neumann & Burks, 1966) reproduction
is a pre-requisite for life. Like the chicken and the egg problem, it can also be
said that life is a pre-requisite for reproduction. Clearly, any decision on this
matter critically depends on the context that is used to define life. If the operator
hierarchy is used, the least complex life form is the prokaryotic cellular operator.
Two arguments currently suggest that life is a pre-requisite for reproduction. The
first states that even though all other organisms originate from the first cell by
reproduction, the first cell itself had an inorganic origin. The emergence of the
first cell thus shows that life does not obligatorily result from reproduction. The
second argument posits that organisms do not need to show reproduction, i.e.,
producing offspring, to comply with the operator-based definition of life; the
operator-based definition demands that organisms show two closures:
autocatalysis and a membrane. Autocatalysis can be regarded as reproduction
On life and death
133
without creating offspring. As Jagers op Akkerhuis (2001) pointed out,
autocatalysis implies that a cell autonomously creates a structural copy of its
information, a process that is called ‘structural (auto-) copying of information’.
Before answering the question of whether the structural (auto-)copying of the
cell‘s information means that it must reproduce, it is important to detail the
concept of information. For the latter, I suggest applying Checkland and Scholes
(1990) definition of information to the autocatalytic set. These authors have
defined information as data with a meaning in a context. In line with this
reasoning, Kauffman (1993) proposed that, by selecting the autocatalytic
process as the context, every catalytic molecule becomes a data-unit with a
catalytic meaning (the ‘purpose’ mentioned by Kauffman 1993, p.388) and
represents a part of the information of the autocatalytic process. Following one
round of autocatalysis, or more rounds to account for the loss of enzymes over
time, the cell contains copies of all of its information. At that moment, it has
autonomously performed structural copying of information and fulfils all the
requirements of the operator hierarchy, even when it does not produce an
offspring. Based on this reasoning, the capacity of autocatalytic maintenance is
an obligatory requirement for cellular life and reproduction is a possible
consequence.
The above implies that it is not relevant for a general definition of life to
distinguish between life forms with or without replication, as Ruiz-Mirazo et al.
(2004) has suggested. The latter authors distinguish ‘proto-life stages’ that do
not show a phenotype-genotype decoupling (soma with genes) from ‘real life’
with genes. In line with the operator hierarchy based definitions, Morales (1998)
warns that “if reproduction is required: This is a troubling development, because
it means that we could not tell whether something is alive unless we also know
that it is the product of Darwinian evolution.” The operator-based definition
considers life as a prerequisite for reproduction instead of reproduction as a
prerequisite for life. Consequently, worker bees, mules, infertile individuals and
other non-reproducing organisms and/or phenotypes are life. This point of view
also solves problems that may arise when demanding that memons be able to
reproduce as a prerequisite for recognising them as life forms. In fact, none of
the cellular memons living today shows reproduction, at least not reproduction of
their neural network structure determining their closure. The things they pass on
during reproduction are the genes of their cells, allowing the development of a
multicellular organism with a neural network, capable of learning but devoid of
inherited neural information other than reflexes.
134
Chapter 5
Life holding its breath
The above chapter shows that reproduction is not a prerequisite of life but a
possible consequence of it. Going one step further, it can also be concluded that
metabolism is not a prerequisite for life. Many taxa such as bacteria, protozoa,
plants, invertebrates and vertebrates have developmental stages showing
natural inactivity (seeds, spores) or reversible inactivation when submitted to
desiccation, frost, oxygen depletion, etc. The inactive state carries the name of
anabiosis, after the process of coming to life again (for a review of ‘viable
lifelessness’ concepts, see Keilin 1959). Another type of reversible inactivity
showing marked similarity with anabiosis is the state of neural inactivity in
memons following anaesthesia. An aesthetic that blocks the transmission of
signals between neurons while leaving the remaining metabolic activity of the
neurons intact causes a reversible absence of neural activity that corresponds to
an anabiotic state of the memon.
Even in the early days of the biological sciences, scholars discussed whether
dried or frozen anabiotic stages are alive at a very slow pace, or whether they
are truly static states of matter. In 1860, the famous Société de Biologie in Paris
wrote a lengthy report on this subject (Broca 1860-1861). Quite importantly, this
report concluded that the potential to revive an anabiotic stage is an inherent
aspect of the organisation of the material of which the object consists and that it
is equally persistent as the molecular state of the matter forming the system. In
short, the Société de Biologie found that “la vie, c’est l’organisation en action”.
Additional support for this conclusion came from Becquerel (1950, 1951) who
subjected anabiotic stages to a temperature 0.01 degree above absolute zero, a
temperature at which no chemical processes can occur, even not very slowly.
Becquerel demonstrated that structure alone is enough to allow revival at normal
temperatures. Anabiosis from absolute zero or complete desiccation has led to
the conclusion that “The concept of life as applied to an organism in the state of
anabiosis (cryptobiosis) becomes synonymous with that of the structure, which
supports all the components of its catalytic systems” (Keilin 1959), or that “life is
a property of matter in a certain structure” (Jeuken 1975). With respect to the
question of: ‘what certain structure?’ the operator hierarchy suggests that all
operators with a complexity similar to or higher than the cell answer this
question.
Life as we do not know it
Considerations about “life as we do not know it” depend on assumptions. As a
context for such assumptions, the operator hierarchy offers two advantages.
First, the operator hierarchy has its basis in the general principle of first-next
On life and death
135
possible closure. Second, the rigid internal structure of the operator hierarchy
offers a unique guide for assumptions about life that we do not yet know.
Based on the general principle of first-next possible closure, the operator
hierarchy shows a strict sequential ranking of the operators. Assuming that
closures act as an absolute constraint on all operator construction, the operator
hierarchy then has universal validity. Support for the latter assumption comes
from the observation that, as far as we know, all operators with a complexity that
is equal to or lower than the molecules seem to have a universal existence. If
this universality extends to the biotic operators, the material organisation of
higher-level operators, such as cells and memons, may then possibly be found
in the entire universe. Such universality would significantly assist in the search
for exobiotic life forms because alien life may show similar organisation to the
life we do know, at least with respect to the first-next possible closures involved.
The demand of closure still leaves a good deal of freedom for the physical
realisation of operators. On other planets, different molecular processes may
form the basis of the autocatalysis and interface of the first cells. Similarly, the
operator hierarchy poses no limits to the actual shape, color, weight, etc. of
exobiotic multicellular organisms. Furthermore, even though the presence of
neural networks may be required for memic organisation throughout the
universe, the operator hierarchy does not restrict the kind of elements producing
these networks, or the details of the neural network structure other than
demanding hypercyclicity and interface.
The rigid internal structure of the operator hierarchy allows predictions about the
construction of life forms that have not yet evolved on Earth. Of course, any
discussion of this subject involves speculation, but the operator hierarchy may
well offer a unique starting point for such a discussion. In an earlier publication
(Jagers op Akkerhuis, 2001), I have indicated various future operator types with
a higher complexity than the cellular hardwired memon. To minimise the aspect
of speculation, I would like to discuss here only the memon immediately above
the cellular hardwired memon (see fig. 5.1), the so-called ‘softwired memon’.
According to the operator hierarchy, this type of memon should be able to copy
information structurally. This means that the organism should be able to copy all
of its information by copying the structure of its neural network. At a lower level
in the hierarchy, cells do this by copying their genetic molecules. Softwired
memons can also do this. They are based on a virtual neural network that
resides in computer memory arrays. During their operation softwired memons
continuously track all their neurons, neural connections, connection strengths
and interactions with the interface. It is therefore only a small step for softwired
memons to read and reproduce all the knowledge in their neural network by
copying these arrays. On these grounds, it may be deduced that softwired
memons (or still higher complexity memons) form the easiest way to satisfy the
demands of the operator hierarchy for the autonomous, structural copying of
136
Chapter 5
information. The operator hierarchy suggests therefore that life as we do not
know it will take the shape of technical memons.
The above reasoning shows that the operator hierarchy offers clear criteria with
respect to different forms of ‘artificial life’. The acceptance of an artificial entity as
life is only possible when it shows all of the required properties of an operator.
Referring to the difference between strong artificial life and weak artificial life,
which do and do not consider a-life entities as genuine life, respectively, it would
be fully in line with the present reasoning to consider as genuine life all a-life
entities that fulfil the requirements for being an operator.
On life and death
Given the present focus on states of matter, it is quite simple to define dead
matter as: all operators that do not fit the general definition of life. It is more
difficult, however, to define death.
Given the current point of view, death represents a state in which an organism
has lost its closure. The use of closure in this definition helps prevent that “….
the properties of an organism as a whole [would be confused] with the properties
of the parts that constitute it” (Morales 1998). However, organisms also loose
their closure during transitions that are part of life cycles and that are not
associated with the organism’s death. For example, the closure of the organism
is lost and a new closure gained when the zygote exchanges its unicellular
organisation for the multicellular state of the embryo and when the multicellular
embryo develops to a memic state. Is it possible to specify the loss of closure
during death in a way that excludes closure losses during life cycles?
With respect to the above question of how to exclude the loss of closure during
transitions in life cycles when defining death, the general process of
deterioration offers a solution. During their lives, organisms deteriorate because
of injury and ageing. The loss of closure marking death is always associated
with the organism‘s irreversible deterioration. Demanding irreversible
deterioration, therefore, helps to prevent that one would be tempted to consider,
for example, a caterpillar as having died, when its tissues are reorganised during
the transition via the pupae to a butterfly. Accordingly, it is possible to describe
death as: the state in which an organism has lost its closure following
irreversible deterioration of its organisation.
Using the above definition, death may occur in either an early or late phase of
the deterioration process, and following the death of multicellulars, a short or
long period may pass until the organism‘s body parts become dead matter. The
latter has its cause in the hierarchical construction of multicellular organisms.
On life and death
137
Accordingly, the loss of the highest closure implies a classification of the
remaining body as an operator showing the first-next lower closure.
Death depends on the loss of closure. To illustrate the contribution of this
statement to the analysis of death, the death of a memon can be used. Due to
the memon‘s strongly integrated organisation, death may occur at various levels
that affect other levels. For example, the multicellular regulation may be the first
to collapse due to the loss of liver functions. After a certain period, this will cause
failure of neural functioning, the latter marking the memon‘s death In another
situation, the neural functions may be lost first, and the memon is the first to die,
dragging its body with it in its fall. However, sometimes enough neural activity
may remain for a vegetative functioning of the memon‘s body as a multicellular
unit. The vegetative state cannot maintain itself autonomously (in principle, a
requirement for a multicellular organism) but it may continue given the right
medical care. If this care is withdrawn, the multicellular body will start
deteriorating after which the cells in the organs and tissues will start dying at
different rates. At a certain point, the multicellular closure is lost, and separately
surviving cells have become the next level operators to die. Physiological
differences between cells now determine the period during which they can
survive in the increasingly hostile habitat of the dead memon, which is cooling
below the normal operating temperature of cells and which shows many adverse
chemical changes such as the lowering of oxygen levels, the release of decay
products of dead cells, etc. Shortly after the memon‘s death, it is possible to take
intact body-parts, organs and cells from its body and sustain their functioning
following transplantation to a favourable environment. For example, the offspring
of cells from the cervix of Henrietta Lane are still cultured as He La cells in many
laboratories.
The inutility of property lists
The above arguments and examples have explored the possibilities of using the
operator hierarchy for creating coherent definitions of life, the organism, living
and death. However, how should the outcome be evaluated? Have the present
attempts led to definitions that could be generally accepted in the field? A way of
evaluating this that has become rather popular is to check the results against
lists of preset criteria. Those who want to evaluate the present approach in this
way may want to examine the following lists of criteria.
Morales (1998) has published a list of properties for a definition of life that
includes the following criteria: 1. Sufficiency (Does the definition separate living
entities from non-living ones?), 2. Common usage (simple classification of easy
examples), 3. Extensibility (Does the definition deal with difficult cases, such as
viruses, mules, fire, Gaia, extraterrestrial life and robots?), 4. Simplicity (few ifs,
138
Chapter 5
buts, ands, etc.) and 5. Objectivity (Criteria are so simple that everyone applies
them with the same result). Emmeche (1997) offers another criteria list for a
definition of life that includes the following: 1. The definition should be general
enough to encompass all possible life forms. (The definition should not only
focus on life as we know it.), 2. It should be coherent with measured facts about
life, (It should not oppose obvious facts.), 3. It should have a conceptual
organising elegance. (It can organise a large part of the field of knowledge within
biology and crystallise our experience with living systems into a clear structure, a
kind of schematic representation that summarizes and gives further structure to
the field.), 4. The definition should be specific enough to distinguish life from
obviously non-living systems. Emmeche (1997) furthermore states that a
definition “should cover the fundamental, general properties of life in the
scientific sense”. Korzeniewski (2005) has also proposed a list of criteria for a
cybernetic definition of life, and Poundstone (1984) has extracted further criteria
for life from the work of von Neumann & Burks (1966). Oliver and Perry (2006)
have suggested a list more or less similar to that of Emmeche (1997) focusing
specifically on properties of a good definition.
With respect to the use of criteria lists, I agree with other authors (Maturana and
Varela 1980, van der Steen 1997) that it is not necessarily an advantage if a
theory performs well or a disadvantage if a theory performs poorly according to a
list of criteria; an approach‘s value does not necessarily correspond to its
performance in these types of checklists. The match depends on the similarity in
major goals and paradigms and the creator‘s influence on the selection and
definition of criteria in a given list. In addition, the selection of ‘favourable’ lists
can lead to false positives.
For the above reasons, I am convinced that it is only possible to evaluate the
currently proposed definitions ‘the hard way’, i.e., by critically examining the
internal consistency and transparency of their logic. In this respect, the present
approach has the advantage of a fundamental bottom-up construction. It starts
with defining elementary building blocks, the operators, and their hierarchical
ranking in the operator hierarchy. To recognise and rank the operators, the
operator hierarchy uses first-next possible closures. In the resulting hierarchy,
the definition of every higher-level operator depends, in an iterative way, on a
lower-level ‘ancestor’ until a lowest-level ancestral system is reached, which is
presumably the group of elementary particles that according to the superstring
theory may have a common basic structure. The result is a strict, coherent and
general framework that is open to falsification: the operator hierarchy.
Subsequently, the operator hierarchy offers a fundament to define a range of
secondary phenomena, such as life, the organism, living and death. Because of
the reference to the operator hierarchy, the present definitions are short, logical
statements that show a high specificity with respect to whether a certain entity
satisfies the definition (list of examples in the following paragraph).
On life and death
139
Testing the definition of life
When using the operator hierarchy as a context for a definition, it is easy to
conclude that viruses, prions, memes or replicating computer programs are not
forms of live. Both a virus with a surrounding mantle and a viral strand of DNA or
RNA are not operators, thus not life. Prions are molecules, thus not life. Memes,
such as texts and melodies, are pieces of coding that memons can decode and
replicate (Dawkins 1976). Accordingly, memes are not operators, thus not life.
Ray (1991) has created computer programs that can replicate themselves onto
free computer space, show mutation, and modify and compete for the available
space in a virtual world called Tierra. Since its start, this virtual ‘ecosystem’ has
seen the evolution of a range of different computer programs. In the same way
as molecular viruses depend on cells, the programs in Tierra depend on a
computer to copy and track their structure. Accordingly, they are not operators,
thus not life. Sims (1994) has used genetic algorithms for evolving virtual
computer creatures with body parts and a neural network with interface. The
simulation of these animal-models allows virtual movement such as finding and
grasping virtual food items. Sims’s programmed creatures may possess
hypercyclic neural networks and on these grounds show similarity to softwired
memons. According to the operator hierarchy, a softwired memon should
autonomously be able to copy its information structurally. Although I am not an
expert in this field, it seems to me that Sims’s organisms do not themselves
keep track of their arrays with information about their interface and neurons,
neural connections, and connection strengths, and that they do not
autonomously organise their maintenance. Assuming that the latter
interpretations are correct, Sims’s computer animals are not yet life.
The use of the present definition also allows the effortless rejection of other
systems that are not operators and sometimes receive the predicate of
‘borderline situations’, such as flames, whirlwinds, crystals, cars, etc. Technical,
computer based memons, however, such as robots, can be operators when they
show the required structure.
To summarise the practical applicability of the present definition of life, I include
a list of the examples that were discussed in the text and supplement them with
some additional cases. The examples in this list form three groups depending on
whether the entities involved are operators or not, and whether they show a
complexity that equals or exceeds that of the cellular operator. In the text below I
use the concept of ‘interaction system’ (e.g. Jagers op Akkerhuis, 2008) for all
systems that are not operators because the interactions of their parts do not
create a first-next possible, new, closure type.
140
Chapter 5
Group A. Systems that are not life because they are not an operator
1. An entire virus particle with external envelope (represents a simple
interaction system)
2. A computer virus based on strings of computer code
3. A flame
4. A tornado
5. A crystal
6. A car
7. A bee colony (The colony is an interaction system, and the bees are
organisms.)
8. A cellular colony not showing the requirements of multicellularity (The
individual cells are organisms and thus represent life.)
9. A colony of physically connected cellular memons (as long as the
individuals lack the required memic closure)
10. A robot (as long as it is a non-memic technical machine)
11. Computer simulations of organisms (including memons) that depend on
external ‘orchestration’
12. A cutting/slip of a plant that cannot potentially show autonomous
maintenance given the right conditions (It lacks the closure required for
multicellularity.)
13. A separate organ, such as a liver or leg (not potentially capable of
autonomous maintenance)
14. Endobiontic bacteria having lost genes that are obligatory for autonomous
maintenance. The transfer to the genome of the host of DNA coding for
enzymes required in autonomous maintenance implies a partitioning of the
autocatalytic closure between the endobiont and its host. Because of this,
the endobiont is no longer an autonomous organism but has become a
special kind of organelle.
Group B. Systems that are operators but that are not life because their
complexity is lower than that of the cellular operator
1. A prion
2. Self-replicating DNA/RNA particles (catalyse their own copying in a solution
containing the right building materials)
3. A DNA or RNA string of a virus that is copied in a cell
Group C. Operators representing life
1. A cutting/slip or other plant part that can potentially maintain itself given
favourable environmental conditions
2. Anabiotic organisms (The fact that they are dried, frozen, etc. does not take
their required closure away.)
On life and death
141
3. Fully anaesthetised animal supported in its functioning by a mechanical
heart-lung support and showing no neural activity (This can be regarded as a
form of memic anabiosis with the potency to become active again.)
4. A computer memon or other technical memon (a memic robot)
5. An artificial cellular operator constructed by humans
6. A exobiotic cellular operator with another chemistry than that found on earth
7. Sterile or otherwise non-reproducing organism (e.g. a mule, worker bee,
sterile individuals)
8. Endoparasites or endosymbiontic unicellular organisms living in cells and still
possessing the full potential of autocatalysis
In conclusion
1. Overviews of the definitions of life from the last 150 years show that no
consensus definition on life exists. In the light of the continuous failure to
reach consensus on this subject, certain scientists have adopted a practical
viewpoint, accepting, for example, the use of property checklists for
identifying living systems. Others have advocated that the need for a
generally accepted definition remains acute. Amongst the proposals for
solving the problem is the suggestion to construct a broader context, a
‘theory of life’ before continuing with attempts to define of life.
2. Inspired by the latter suggestion, the present paper invokes a classification of
the generalised particle concept, called the ‘operator hierarchy’. This
hierarchy has several advantages for defining life: first, it offers a general
context for including and differentiating between life and non-life, and second,
it offers the unique possibility to extrapolate existing trends in the evolution of
material complexity and to use these as a guide for discussions about ‘life as
we do not know it’.
3. In close association with the reviewed literature, the use of the operator
hierarchy allowed the following definitions to be suggested:
A. From the viewpoint of the evolution of material complexity, life is: matter
with the configuration of an operator, and that possesses a complexity
equal to or even higher than the cellular operator.
B. Living describes the dynamics of an operator that satisfies the definition
of life.
C. The definition of unitary organisms can take the form of: the operators
that fit the definition of life.
D. A multicellular organism (the cellular operator showing the multi-state)
is: a construction of mutually adhering cells showing obligatorily
recurrent interactions based on the same interaction type, that has the
potential of maintaining its functioning as a unit and that does not show
memic structure
E. Dead matter applies to all operators that do not fit the definition of life.
142
Chapter 5
F. Death is: the state in which an organism has lost its closure following
irreversible deterioration of its organisation.
4. From the discussion of examples in the literature, it was concluded that the
present set of definitions easily distinguishes life and non-life regardless of
whether this is tested using the ‘obvious examples’, the ‘borderline cases’ or
‘life as we do not know it’. This suggests that the present approach may well
offer a practical step forward on the path towards a consensus definition for
the states of matter representing ‘life’.
Acknowledgements
The author would like to thank Herbert Prins, Henk Siepel, Hans-Peter
Koelewijn, Rob Hengeveld, Dick Melman, Leen Moraal, Ton Stumpel and Albert
Ballast for constructive discussion and/or comments on the manuscript and
Peter Griffith for English correction.
On life and death
143
The issue of ‘closure’ in Jagers op Akkerhuis's operator
theory
Comments by Nico van Straalen
In this issue of Foundations of Science Jagers op Akkerhuis (2010a) proposes a
definition of life based on his earlier theory of operators. A great variety of
objects fall into the category of operator, and by introducing this term Jagers op
Akkerhuis was able to draw a parallel between elementary particles, molecules,
cells and multicellular organisms. The common denominator of these operators
is their autonomous activity and maintenance of a specific structure.
Consequently, operators were classified in a logical and hierarchical system
which emphasizes the commonalities across what is normally called non-life
(atoms, molecules) and life (cells, organisms). One very attractive aspect of the
classification is that it joins the objects traditionally studied by physicists,
chemists and biologists into one overarching system. Obviously, the hierarchy
crosses the traditional border between life and non-life, so it should be possible
to develop a definition of life from the operator theory. This is what Jagers op
Akkerhuis attempts to do in the present paper. However, I believe he misses the
point.
In the operator hierarchy, successive levels of complexity are separated by
‘closure events’, e.g. when going from hadrons to atoms, from molecules to cells
and from multicellular eukaryotes to memic organisms. One of these closure
events actually defines the origin of life: the transition from molecules to cells.
Death, as defined by Jagers op Akkerhuis, is the loss of this closure, a fall-back
from cells to molecules. There is another important transition, the origin of self
consciousness, a closure event that accompanies the highest level of complexity
in the classification of operators. Life with this level of complexity (maybe call it
‘hyper-life’?) is included in Jagers op Akkerhuis's definition of life.
Another interesting aspect of the operator system is that it is strictly hierarchical,
that is, every operator can be classified on a more or less linear scale and the
big leaps forward are punctuated by closures on that scale. This aspect of the
system is reminiscent of the ‘Great Chain of Being’, or scala naturae, which was
the dominating view of life for many centuries. In evolutionary biology, it is now
recognised that pathways can split and run in parallel, maybe even achieving
similar closures independently from each other. I am not sure how this aspect
fits into the operator classification of Jagers op Akkerhuis.
To define life in terms of the operator theory I believe the focus should be on the
transition from molecules to cells and the closure aspects of this event. In other
words, the closure of operating systems defines life better than the classification
of operators. However, Jagers op Akkerhuis seems to add another seemingly
144
Chapter 5
hopeless definition of life to the nearly 100 already existing. Classifying what is
life and what is not is, I believe, a rather trivial exercise. Everybody knows that a
flame is not life, and it only becomes a problem when you spend too many
words on it. Rather than classifying things into living and non-living entities I
believe the challenge is to understand how the transition from non-life to life can
take place, that is the how the closure in Jagers op Akkerhuis's hierarchical
classification of operators, comes about.
The issue of closure is intimately linked to that of emergence. Both concepts
recognise that the characteristic properties of a living system cannot be reduced
to its component parts only, but also depend on the way in which the
components are organised in a network. The properties that arise from
interactions between components are said to be ‘emergent’. Emergent
properties are not shared by the components; they ‘appear’ when many
components start interacting in a sufficiently complex way.
The concept of emergence plays an important role in genomics, the science that
studies the structure and function of a genome (Van Straalen & Roelofs, 2006).
After about a decade of genome sequencing, scientists started to realise that the
genome sequence itself does not define the organism. The human genome
turned out to contain no more than 24.000 genes, much less than the earlier
assumed 124.000. This raised the question how it could be possible that such a
complicated organism as a human being could be built with so few genes.
Obviously the pattern of gene and protein interaction defines human nature
much more than the genes and proteins themselves. A new branch of biology
was defined, systems biology, which was specifically geared towards the
analysis of interacting networks, using mathematical models (Ideker et al.,
2000).
Schrödinger (1944), in discussing the question ‘What is life?’ foresaw a new
principle, not alien to physics, but based on physical laws, or a new type of
physical laws, prevailing inside the organism. These are the kind of laws that
systems biology is after. The operator classification of Jagers op Akkerhuis is an
important step because it emphasizes the continuity between physical systems
and biological systems. However, the challenge of defining life is not in
classification but in understanding the closure phenomenon by which life
emerged from non-life.
On life and death
145
Definitions of life are not only unnecessary, but they can do
harm to understanding
Comments by Rob Hengeveld
In his paper, Jagers op Akkerhuis (this volume) refers to a list of almost 100
different definitions subsequently having been given in the literature to the
phenomenon of life as we know it. These definitions may even have a more
general application or meaning than that concerning life on earth only. That is,
also to some form of life as we don’t know it, even though we don’t know it. Like
other authors, he feels that all this activity messed things up. Thus, Jagers op
Akkerhuis mentions authors emphasising “the seeming hopelessness of the
situation”, some of them adopting “a pragmatic/pessimistic viewpoint”. Others
would have suggested “that life cannot be defined, that its definition is not
important for biology”, or that “a comprehensive definition of life is too general
and of little scientific use”. Finally, only “living processes may be defined” which
“cannot be reduced to physics”.
His theory based on the criterion of hierarchically arranged operators would tidy
up this mess a little. I feel, though, that the introduction of his own definition “life
may be regarded as a special realisation of the evolution of material complexity”
brings the 98 existing definitions even closer to 100. Worse, this theory and
definition will confuse our biological issues even more by their circularity of
reasoning. They are circular because his operator concept “emphasizes the
autonomous activity of the entities involved, which ‘operate’ in a given
environment without losing their individual organisation”. How do we distinguish
the autonomy of processes in early living systems or even in present-day
molecular biological ones from those of non-living processes? Also, activity,
operation, and organisation are concepts connected with living systems and
their functioning. Furthermore, individual organisation smells of one of the
criteria on which some earlier definitions have been based. Thus, recognising
something as living depends on criteria derived from known, recent living
systems; a bean is a bean because it is bean-shaped.
When, as a beginning ecologist, I was studying ground beetles, and later as a
biogeographer, I never felt any need for a definition of life. Then, such a
definition was clearly useless. More recently, being concerned with questions
about the origin of life, that is concerned with processes ultimately having
resulted in a beetle as a living system, I came to realise that most, if not all, of
these definitions were designed particularly within this context of the origination
of living systems. However, we don’t need to define the moon to understand its
origin either. Yet, they not only seem useless, they are even harmful. Adopting
certain criteria on which to base the one or the other definition, authors easily
force themselves to look into the wrong direction. Or even at the wrong
146
Chapter 5
biogenetic phase, too late in the development of life. For example, not only is
‘organisation’ difficult to delineate objectively at a molecular level, and this
without circularity, but depending on a subjectively chosen threshold level, it
easily excludes initial phases from analysis, however significant these could
have been. Continuing along such a misdirected road is fruitless.
Thus, the criteria used, such as a certain level of organisation, are always
derived from present-day life forms, from processes or structures that may not
have existed in the early biogenetic phases. One criterion, as one of many
examples, points at macromolecules, although these will have developed from
earlier oligomers (see, for a clear example, Eck and Dayhoff, 1966). Another,
widely applied criterion derives from the present prevalence of carbon as a
principal biochemical constituent. Yet, carbon forms very stable molecules, as
do its neighbours in the Periodic table, nitrogen and oxygen, for example. They
are difficult both to form as well as to break down again, which is therefore
usually done by enzymes. These enzymes, plus the enzymatic apparatus they
together form, must have developed earlier, before carbon could have been
taken on board biochemically (see Hengeveld and Fedonkin, 2007). Moreover,
in their turn, individual enzymes are often very complex macromolecules, which
not only must have been derived evolutionarily from more primitive ones, but
they have to be formed by and operate within an intricate biochemical apparatus
in which DNA is pivotal. Yet, DNA itself requires the operation of a very complex
system of repair enzymes, etc., plus the mediation of spliceosomes and
ribosomes for the final construction of those enzymatic macromolecules. Clearly,
carbon as an element must have been inserted into the biochemistry only at a
later, evolutionarily more highly developed stage of biogenesis.
Personally, I prefer to abstain from using definitions in this context. This differs
from asking what requirement is needed to form a molecular bond, of a system
of molecules, etc., any form of organisation, biological or non-biological. This
puts the problem within the thermodynamic realm. A basic requirement, one that
can be met by several properties, therefore differs from a property, physical,
chemical, biological, or socio-economic; instead, it defines both the process and
the shape of molecules taking part in it (Hengeveld, 2007). It defines the
properties. It’s the resulting processes happening and developing which are of
interest, for the understanding of which a definition of life is irrelevant. It does not
add anything.
Formulating the study of biogenesis in terms of processes happening and
developing precludes the design of definitions, which are more likely to be
applied to static or stable situations. And which are, already for that reason only,
to be shunned. Defining life is not a part of our scientific endeavour.
On life and death
147
Explaining the origin of life is not enough for a definition of
life
Reaction of Gerard Jagers op Akkerhuis to the comments and questions of
Rob Hengeveld and Nico van Straalen.
I thank the commentators for their reactions, both positive and negative, giving
me the opportunity to elucidate some important aspects of the presented theory.
As Van Straalen indicates, the operator hierarchy offers valuable innovations: Firstly,
the hierarchy ‘… joins the objects traditionally studied by physicists, chemists and
biologists into one overarching system.’ Secondly, ‘...it is strictly hierarchical’. I think
that precisely these two aspects make the operator hierarchy a unique tool for
defining life in a way that simultaneously addresses all the different organisational
levels of living entities, e.g. prokaryotic cells, eukaryotic cells, pro- and eukaryotic
multicellulars and neural network organisms, including future ones based on
technical neural networks.
Further reactions of the commentators indicate that, probably due to the novelty of
the operator theory, certain aspects require further explanation. I will discuss some
essentialities in the following lines.
Hengeveld criticizes an asserted circularity in reasoning, in the sense that living
operators are defined by means of concepts, which are derived from living systems.
The confusion on this point results from my explanation in the paper. There I
indicate that the name operator originates from the operating (in a very general
sense) of individual entities. It may be reassuring to Hengeveld that the origin of the
name ‘operator’ shows no direct relationship with the definition of the operators as
system types. The entire set of all operators is defined as follows: based on the
presumed existence of a lowest complexity operator, every system that belongs to
the operator hierarchy resides at exactly one higher closure level than its precedinglevel operator. Every closure level is defined by the occurrence of one or two firstnext possible closures. Although this is a recursive definition in the sense that every
operator in principle depends on its preceding-level operator, its hierarchical
architecture precludes circularity of reasoning.
Hengeveld furthermore states in a general way that definitions of life ‘are always
derived from present day life forms, from processes or structures that may not have
existed in the early biogenetic phase’. This general criticism does not apply to the
operator hierarchy. The reason is that both abiotic and biotic operators are all
defined using first-next possible closure. In fact, the operator theory turns the
argumentation of Hengeveld upside down, hypothesising that limited possibilities for
reaching first-next possible closure have acted as a blue-print for the essential
construction properties we recognise in abiotic elements and organisms.
148
Chapter 5
I agree wholeheartedly with both Hengeveld’s and Van Straalen’s argumentation
that we need to increase our understanding of the processes that have caused
life. I strongly support the search for bootstrapping mechanisms allowing simple
system states/elements to autonomously create more complex system
states/elements (e.g. Conrad 1982, Martin & Russell, 2003, Hengeveld, 2007).
In fact, every closure step in the operator hierarchy is the product of a specific
(the first-next possible) bootstrapping mechanism. With respect to specifying the
closure types resulting from such bootstrapping mechanisms, I consider concise
and general definitions as indispensible tools, being helpful (instead of harmful!)
in our search for the essences of the evolution of matter. Thus when Hengeveld
advocates that he prefers ’… to abstain from using definitions in this context.’ I
find his viewpoint surprising for two reasons. The first reason is that even a very
thorough understanding of specific reaction mechanisms will not automatically
result in a general definition of a meta-aspect such as the type of material
organisation defining living entities. The second reason is that I think that
accurate definitions are simply a way to improve the precision and
communication of science: sloppy definitions lead to the development of sloppy
theory and a lack of definitions leads to no science at all.
Referring to a demand for a mechanistic focus when defining life, Van Straalen
states that ‘the challenge is to understand how the transition from non-life to life
can take place’ as this can explain how the classification of operators comes
about. Also in his last sentence Van Straalen writes that ‘…, the challenge of
defining life is not in classification but in understanding the closure phenomenon
by which life emerged from non-life’. Both statements being true, it is
nevertheless impossible to construct an overarching theory such as the operator
hierarchy if one limits his view to the mechanisms explaining one single step
involved.
The warm interest of Hengeveld and Van Straalen for mechanisms that could
explain the origin of life is understandable, because it frustrates the scientific
community that science is not yet able to de novo synthesise life, not even in the
form of a primitive cell. This general focus on the construction of life seems,
however, to have caused a tunnel vision with respect the definition of life.
Imagine that we would be able to explain the cell, and even construct it, would
this then mean that we would have a proper definition of life in all its forms,
including multicellular organisms and neural network organisms? The answer is
a clear NO. If everything that is based on living cells would be life, then a donor
organ and a fresh, raw steak would also be life. Moreover, any technical being,
however intelligent, could never be called life, because it is not based on cells.
This proves that a focus on cells alone is not enough. We need to broaden the
scope and define all levels of organisation associated with higher forms of life. It
is my personal conviction that, for the latter goal, the operator hierarchy offers a
unique and unprecedented tool.
Getting the biggest part of the pie
Chapter
6
Getting the biggest part of the pie
Using resource dominance to explain and
predict evolutionary success
G.A.J.M. Jagers op Akkerhuis and C. Damgaard
Oikos 87:609-614, 1999
“Without competitors there would be no need for strategy.”
(Keniche Ohnae)
149
150
Chapter 6
Abstract
In this century, evolutionary biology has been successful in using fitness as a
measure for evolutionary success. Although fitness has the intuitive association
of "something which is better", a restricted definition in the form of "the number
of successful offspring" is often used in biology. This definition is convenient
because it is, in principle, easy to count offspring. Yet, the explanatory and
predictive capacities of such fitness measurements are restricted and the
method is restricted in its use to reproducing entities.
To better understand traditional fitness measurements, we suggest that more
attention be paid to underlying factors. For this purpose we advocate to focus on
"resource dominance". The idea is that any unit of evolution which is good at
dominating the resources needed for maintenance, growth and/or reproduction
will also show a high evolutionary success. We show that high resource
dominance can be reached via different strategies. Additionally, we discuss how
the capacity for resource dominance is related to the complexity and efficiency
of the evolving unit.
Getting the biggest part of the pie
151
Introduction
Ever since Darwin (Darwin 1859) and Wallace (Wallace 1855), evolution has
been expressed as a struggle for life and survival of the fittest. Despite the
general intuitive interpretation of a fit entity as being better in some respect, a
more specific use of fitness has become generally accepted in biology (Frank
1997, Gavrilets 1997, Fear and Price 1998). This states that the fitness of an
entity is proportional to the number of successful offspring. Using this definition,
fitness is a convenient measure for evolutionary success, because it is, in
principle, straightforward to measure the number of offspring. However, there is
little predictive power in such a concept of fitness, and it is left open when it may
be judged that an offspring has been successful. Additionally, fitness in this
definition cannot be used as a criterion for the evolutionary success of units
during periods in which they do not reproduce.
To improve on this situation we advocate the studying of the underlying factors
which are responsible for the number and the quality of offspring. This, we claim,
is done best by focusing on the capacity of organisms to gain dominance over
resources, which we term their resource dominance.
In this study we will discuss different aspects of resource dominance as the
basis for evolutionary success. We will present several important resource
dominance strategies and pay attention to the interaction between resource
dominance and the complexity of the evolving unit. By focusing on resource
dominance it will become apparent that knowledge of the life history and
ecological strategies of the different entities is critical in understanding the
evolutionary processes.
To allow a broad covering of the resource use strategies and complexity of
evolving units, we will apply the units of evolution approach. This has been used
by several authors (Hamilton 1964, Lewontin 1970, Hull 1980) as a way to study
evolution of units in relation to their position in the organisation of a system.
Examples of such units and levels are the DNA in the cellular environment,
mitochondria in the eukaryotic environment, endosymbiontic eukaryotes in other
eukaryotes such as has been observed for example in the protist Peridinium
balticum (Bardele 1997), individuals in populations, and groups of individuals,
i.e. social insects in ecosystems.
By focusing on resource dominance as a criterion for evolutionary success, the
concept can be applied to entities which evolve their structure without showing
reproduction, such as companies and evolving neural networks. Additionally,
evolutionary processes in different entities can more easily be compared.
152
Chapter 6
Figure 6.1. A dimensional comparison of resource space, resource dominance space,
and fitness. The first part shows the space of all elements which, in principle, could act as
a resource for evolving entities. The thick arrows in the second part show those parts of
this space which are actually dominated by a particular evolving entity. The last part
shows how the fitness of an entity as a measure for its evolutionary success can be
obtained from resource dominance space via a fitness function. Note that the parameter
reduction involved in obtaining fitness cannot be inverted to obtain information about the
resource dominance strategies laying at the basis of fitness.
Resource dominance
The concept of resource dominance is based on the assumption that evolutionary
success depends on the capacity to gain equal or larger dominance over limiting
resources than competitors. Everything else being equal, a plant gaining access to
the most light will grow fastest, and a small animal will lose more fights.
The focus on resources in the present paper is based on the necessity for
organisms to use resources for their proper functioning, which minimally requires
their maintenance. Under more favourable conditions resources can be
converted into more biomass, leading to growth and/or reproduction.
Fundamental demands for the use of resources during all these processes stem
from the out-of-equilibrium state of organisms as dissipative systems which only
can maintain and/or increase their internal organisation at the costs of entropy
increase in the surrounding environment (Prigogine and Stengers 1984). The
recognition that demands are fundamental to organisms forms, for example, the
core of the ‘supply-demand’-based models of Gutierrez (Gutierrez 1996,
Schreiber and Gutierrez 1998).
Attempting to define resources in accordance with the above, we propose the
following definition: resources are all entities that organisms, alone or as a
group, can use for their functioning. This definition allows much freedom. It
includes both consumables, which are degraded when used, and what could be
considered as durables, for which the use does not markedly affect the state. It
Getting the biggest part of the pie
153
also includes the information, the ideas, communicated in social interactions as
shown by people and some animals.
As the second part of the resource dominance concept we have given
preference to dominance over use, because the dominance by evolving units
over resources may extend beyond the use. Examples of such dominance are
allelopathy and territory.
Quantification of resource dominance
A measure for resource dominance of an individual or population has to be
multidimensional with as many dimensions as there are different types of
resources that play an essential role in the evolutionary success of the particular
individual or population (Fig. 6.1). For each resource dimension the degree of
dominance of the evolutionary unit has to be measured. The resulting measure
can be compared with other evolutionary units that are competing for some or all
aspects of the resource space. If two organisms compete for a common
resource, for example space, their competitive ability may to a large extent
depend on the access to non-overlapping, or ‘external’ resources. The more
external resources can be used to increase their competitive ability, the higher
their competitive capacity for any common resource will be.
The evolutionary success of a unit can now be measured directly as the
increase in the degree of dominance over the resource space, but it is
convenient to reduce the multidimensional resource dominance space to the one
dimensional measure of evolutionary success: fitness (Fig. 6.1). The
understanding of the function that reduces the multidimensional resource
dominance space into the one-dimensional fitness measure will give valuable
information about the underlying evolutionary processes, i.e. how much does the
dominance of a specific resource contribute to the evolutionary success of an
evolutionary unit.
154
Chapter 6
Figure 6.2. Different strategies for organisms and groups of organisms to increase their
resource dominance. The physical size of the organism is shown as the inner, shaded
circle. The outer, dotted circle represents the amount of dominated resources.
Getting the biggest part of the pie
155
Resource dominance strategies
When talking about resource dominance, the question can be asked which
strategies evolving entities can follow to increase this property. Some strategies
are summarised in Fig. 6.2. We have ranked the options roughly in order of
simplicity. The examples allow increased dominance through: 1) internal reallocation, 2) larger size, 3) internal re-organisation and adaptation to new
resources, 4) co-operative relationships, and 5) reproduction.
The first strategy in Fig. 6.2 refers to an increase in resource dominance by a
‘smarter’ or more aggressive strategy aiming at the same ‘tasks’, whilst the
qualitative aspects of the internal organisation of the unit remain unchanged.
This implies the reallocation of resources by means of phenotypic plasticity to
expand certain existing internal/external processes. Examples of situations
where this applies are internal reorganisation leading to induced production of
an antibiotic chemical or to hormonal regulation of territorial aggression.
The second strategy realises an increase in dominance simply by an increase in
size. In competitive environments a plant that is twice the size of another plant
may grow more than twice as fast (asymmetric competition, Weiner 1990).
Bigger animals win more fights. Larger brains, of the same quality, have a larger
thinking capacity. Yet, size poses also evolutionary limits. The size of organisms
puts increasing limits on internal transportation and large animals have to deal
with small surface-to-volume ratios, with many consequences for physiology,
food intake and mobility.
The third strategy is based on innovation, with which we refer to the gaining of
access to new resources. This requires the incidental or deliberate development
of new internal organisation. Examples of innovative events during evolution are
the shift in metabolism that certain bacteria have shown from anoxybiontic
organotrophs to phototrophs, the shift from multicellular life to neural networks,
which allowed the emergence of co-ordinated mobility based on sensory inputs,
and the shift from aquatic to terrestrial life forms. The first organisms possessing the
above properties got access to a massive resource of sunlight to ‘fuel’ their
chemistry, a much better escape or predation behaviour and the capacity to learn
from experiences, and a large space for invasion with little competition, respectively.
The fourth strategy involves co-operation. It includes all situations in which any
form of living together leads to an overall gain in resource dominance. The
evolutionary success of such co-operations is convincing (Herre et al. 1999).
Most markedly, this is illustrated by the presence of mitochondria in the cells of
fungi and animals, and the presence of mitochondria and chloroplasts in the
algae and the plants (Blackstone 1995). That mitochondria and chloroplasts are
156
Chapter 6
symbiotic cells and not just organelles was first proposed by Margulis (1979) in
the form of the ‘endosymbiont hypothesis’, which is now generally accepted. In
addition to endosymbiontic symbiosis, there exist many other forms of cooperation between cells and/or organisms which are important in evolution
(Margulis 1998, Michod 1999). These include symbionts dwelling in multicellular
organisms, colonies of physically linked individuals such as in some corals, and
colonies and other forms of symbiosis of freely moving individuals of the same or
of different species. Behavioural co-operation may involve co-ordinated
behaviour and the division of labour, for example during hunting and other
activities by groups of social organisms.
The fifth strategy, reproduction, is an important characteristic of biological life,
because it forms the basis for the ‘tree of life’, which represents the phylogenetic
relationships connecting present organisms to past ancestors. Evolutionary units
have been selected in different environments for different strategies of
reproduction which locally assure a maximum number of offspring with high
resource dominance, and consequently the highest evolutionary success.
Annual plants and some fish reproduce only once, whereas other evolutionary
units reproduce some or many times over a time span. Some produce many
small offspring, which may be widely dispersed but have a small probability to
obtain a similar resource dominance position as their parents; others
concentrate resources and parental care in a few offspring with a larger
probability of being successful.
Evolution will select for the best combination of strategies in a
given environment
The different resource dominance strategies can be regarded as an evolutionary
‘toolbox’. Organisms have access to different combinations of strategies for
different circumstances. Not every tool is appropriate for every situation. Even
though reproduction may be the strongest resource dominance strategy in the
long term, every organism may experience periods during which it is difficult to
raise young, and other resource dominance strategies are temporarily more
rewarding. Also the use of size to increase resource dominance depends on
circumstances. In competitive situations a large size is mostly an advantage. But
for algae that are cultured in a stirred chemostat of which half the volume is
replaced by fresh medium every 20 minutes, the situation is different. If their
generation time equals 20 minutes, the algae have a chance of 50% that one of
the offspring remains in the chemostat. This will maintain the culture. An
increase in size, to increase resource dominance, will not work, because a large
size generally requires a longer reproductive period, which dilutes the large
algae from the medium. Instead, rapidly reproducing, small cells form the best
dominance strategy.
Getting the biggest part of the pie
157
Complexity and efficiency
When discussing resource dominance and resource dominance strategies, it is
important to realise that organisms only can show resource dominance because
they have a certain construction, which allows them to be active in the world.
Resource dominance, therefore, depends on construction aspects of the
organism. This is the reason that we advocate the integration of resource
dominance theory with an analysis of major structuro-functional aspects of
organisms. As will be explained presently, we have selected complexity and
efficiency as the main aspects of the relationship between resource dominance
and construction.
Complexity
There exist many viewpoints on complexity (Horgan 1995, Kot et al. 1988,
McShea 1991, Jagers op Akkerhuis and van Straalen 1999, Sole et al. 1999).
Most of these refer to the patterning in dynamic interactions or to a hierarchical
layering of structural features and related dynamics. Here, however, we analyse
complexity from a structuro-functional viewpoint, defining it as any form of
internal and external functional diversity as can be supported by the internal
organisation of an evolutionary unit. The overall complexity of an organism thus
depends on the presence of specific physiology, eyes, limbs, etc. More in detail,
the contribution to the organism's complexity of these different phenotypic
aspects can be analysed by means of a multivariate summation. An eye may, for
example, be judged for its capacity to observe different wavelengths, its
resolution, e.g. pixel size, its speed of accommodation, its light sensitivity, etc.
Note that by using the above definition of structuro-functional complexity, we
deliberately avoid discussions about the exact internal structure being
responsible for a certain functional diversity. The reason is that both a large,
badly organised structure and a minute, well-integrated structure may allow an
entity to show the same functional diversity. Accordingly, the above definition
recognises them as showing the same complexity. Yet, these systems will
almost certainly differ in efficiency.
Efficiency
Efficiency is regarded here as the lowest resource use, when comparing two
processes leading to a functionally similar product. Efficiency is an important
aspect of resource dominance because it improves the performance of all
resource dominance strategies. A high efficiency allows a larger dominance at
the same size and a more economic use of the same resource for maintenance,
growth and/or reproduction. Additionally, a high efficiency allows more
organisational excess that may lead to innovations. Efficiency plays also a role
in co-operation. New interactions involved in cooperation generally have their
costs, but can cause a net profit typically because of the gain in overall efficiency
158
Chapter 6
and resource dominance. A special aspect of efficiency is buffering, with which
we refer to the creation of internal or external stores of certain resources when
these are plenty, to be used in periods of shortage. When the times are hard, the
evolutionary units showing the highest efficiency in their buffering will be able to
continue their normal functioning the longest. The others will be less good
competitors and experience a selective disadvantage.
We regard efficiency as logically independent from complexity and related
resource dominance strategies. The reason is that efficiency can be an aspect of
each of the different strategies. There are many situations in nature, however, in
which complexity and efficiency occur in close association. For example, cooperation of entities will frequently increase the functional properties in close
association with the internal complexity increase.
The close bonds between resource dominance, complexity and efficiency have
to be integrated in evolutionary theory in order to explain, i.e., the competitive
advantage of the ‘fittest’ entities during selection processes, and, over many
generations, the apparent increase of complexity during evolutionary time in
specific branches of the phylogenetic tree.
Discussion
Above we have presented a multivariate analysis of evolutionary success based
on resource dominance, complexity and efficiency. The use of these factors for
measuring evolutionary success makes a general approach possible, which
includes reproducing entities as well as non-reproducing entities. This implies,
however, that the traditional focus on fitness has to be accompanied by an
analysis of the underlying resource space, in order to develop explanatory and
predictive models for evolutionary success.
A special aspect of the present approach is that it integrates resource
dominance and the complexity of the evolving units, either directly because high
resource dominance in a competitive environment requires a high complexity, or
indirectly when the forces of selection act against inefficiency. Yet, it remains the
broad spectrum of environmental forces which determines in which direction the
complexity of a unit can evolve.
At the one end of this spectrum we find the situation where competition is
absent. This applies to environments in which resources are limited due to
abiotic causes, whilst the organism experiences a random chance of mortality.
An example hereof is the above discussed well-stirred chemostat with algae.
The highest resource dominance in these situations is reached by the organism
Getting the biggest part of the pie
159
reproducing fastest. It will be clear that this requirement acts as a severe
constraint on complexity increase.
The other end of the spectrum is formed by situations in which the environment
is experienced as relatively stable, or stable in its variability, and in which the
availability of resources as well as mortality are mainly determined by
competitive interactions. As the energy that an organism has to spend on
competitive interactions is correlated with the frequency and extent of contacts,
all competition based approaches are necessarily density dependent (Witting
1997, Heino et al. 1998). In a competitive environment the combination of a
lower complexity, in the sense of less phenotypic possibilities, and lower
efficiency necessarily causes lower resource dominance, and will be selected
against. Accordingly, evolutionary disadvantage can only be avoided by any
combination of changes in complexity and efficiency yielding a net increase in
resource dominance of a unit. This leaves still much freedom to different
strategies, such as ‘specialists’, which may trade complexity for efficiency, or
‘generalists’, which do not care much about efficiency in some aspects. It also
allows for the coexistence within a population of different strategies for resource
dominance. An example is the coexistence of dominant male deer with fullgrown antlers, and male deer without antlers mating in secret with the female
deer of a herd.
Although it may not be the most likely to occur, an especially favourable
combination in the light of selection is that of complexity increase in combination
with efficiency increase.
Rates of improvement
According to the above reasoning we can regard evolving units as being
selected for a combination of complexity and efficiency which allows for a
combination of resource dominance strategies which, in a given environment,
yields the highest increase in resource dominance. Their success in these
combined efforts, e.g. their evolutionary success, can be measured either in
terms of their present resource dominance or in terms of the increase and the
acceleration hereof. Focus on these different rates is of marked importance,
because those which are best now may soon lose to those which get better all
the time, whilst both these strategies will be beaten in the long term by units
which can increase the rate of their improvement. The ecological importance of
the acceleration of improvement has been stressed by Wagner and Altenberg
(1996), who have asked attention for the evolution of evolvability.
Finally, we consider it a theoretical advantage that the resource dominance
approach makes in principle no assumptions about the kind or combination of
160
Chapter 6
resources which are used, about the kind of units of evolution involved, or about
the combination of resource dominance strategies that is followed. Therefore,
the competition of animals and plants for nutrients, light, etc., can be analysed in
the same way as competition between companies on the basis of money, power,
etc. This offers a common basis for the study of evolutionary processes in
different disciplines, such as biology and economy, which is not offered when
using fitness as the evolutionary measure. We consider this an exciting
possibility for future study.
Acknowledgements
This study was performed within the framework of the programme of the Danish
National Environmental Research Institute (NERI): `Biotechnology: elements in
environmental risk assessment of genetically modified plants'.
Chapter
Definitions behind the operator theory
“Make everything as simple as possible, but not simpler.”
(Albert Einstein)
7
162
Chapter 7
The operator hierarchy can be represented as a simple scheme. However, the
reasoning behind this scheme is intricate and, involves many different steps and
definitions. Because transparency is a major asset in science, the aim of this
chapter is to explain and define the basic aspects of the operator hierarchy.
First, some general aspects are defined, such as a ‘system’, a ‘type’ and a
‘closure’. These general definitions form the basis for higher level definitions,
e.g., the ‘operator’ and the ‘operator hierarchy’. The definition process is not a
goal in itself but is meant to sharpen the reasoning involved in the operator
theory and to prevent confusion in situations where the operator theory uses
certain concepts in a way different from other theories.
Most fundamental to the operator theory is the notion of a system itself. The
word system dates back to ancient times and is derived from the Greek word
sunìstanai, which means ‘together establishing’. Another fundamental concept
to the operator hierarchy is that of ‘type’. The use of types is helpful when
referring to general properties of a group of systems without bothering about the
organisational particulars of every individual system involved.
System
The system concept relates to entities that can be grouped by functionality
because of past, present or future interactions and/or relationships. The entities
that, by their relationships, define the group are considered the system’s
elements. Frequently, the elements and their interactions/relationships define,
include or relate to some kind of system limit that reduces the number of
elements being considered. With respect to natural systems, the system concept
is limited to the universe or a subset of it. The latter includes thoughts and
abstract concepts represented by oscillation states in networks of interacting
neurons. Depending on the type of entities and interactions, different systems
can be recognised, such as ecosystems, cybernetic systems, physical particles,
welfare systems, logical systems, organisms, transportation systems, taxonomic
systems, populations, systems for storing and retrieving objects or data,
economic systems, etc.
The viewpoint that systems consist of elements and interactions is well covered
by Von Bertalanffy (1950, 1968). Von Bertalanffy defines systems as ‘a complex
of interacting elements’ where ‘interaction means that the elements stand in a
certain relation’. Hall and Day (1977) have expressed similar viewpoints. They
state that ‘Any phenomenon, either structural or functional, having at least two
separable components and some interaction between these components may
be considered a system’. De Wit (1978) describes a system as ‘a limited part of
reality, the elements of which show a certain relationship’. Mesarovic and
Takahara (1975) propose an abstract interpretation as they define a system as a
subset of the set of all possible input-output relations. Heylighen (1995)
suggests another approach and defines a system as a constraint on variety,
Towards consistent definitions
163
where the constraint determines the invariant identity -- which allows us to
distinguish states that belong to the system’s limits -- in spite of a variety of
appearances of the system.
In 1865, Claude Bernard indicated that defining a system represents a mental
undertaking when he stated that ‘Les systèmes ne sont pas dans la nature, mais
dans l`esprit des hommes’. In other words, nature knows only phenomena while
systems are our mental abstractions thereof.
Type
A type is used to group entities by a common property. Using types of types
creates a type-hierarchy in which lower-level types show commonalities that can
be grouped into higher-level types. For example, the higher-plant-type and the
animal-type can be grouped within the multicellular-organism-type. In
mathematical type theory, there exists a lowest type, the elements of which have
no members. Such elements correspond to the ‘ur-elements’ of certain set
theories. The following types are frequently used with the operator hierarchy:
 Process type: a comparable aspect that can be recognised in the dynamics
of different entities. For example, catalytic activity is a common property of all
enzymes, regardless of the specific reaction that they catalyze.
 System type (general): this type of a system is defined by comparable
aspects with respect to (1) the rules for selecting elements, (2) the properties
defining the elements and (3) the relationships that may exist between the
elements.
 Closure type: a closure type is defined by the presence of comparable
aspects with respect to the element and mechanism types causing the
closure. Departing from the same system type, different self-organising
processes may lead to different closure types. For example, if a bacterial cell
is chosen as the start, one route leads to the closure type of the
endosymbiont/nucleus while another route leads to the closure type of the
multicellular.
Closing in on first-next possible closure
In addition to systems and types, two other concepts play an important role in
the operator theory, namely that of a ‘first-next possibility’ and that of ‘closure’.
In this study the definition process of a first-next possibility was started with a
literature review that led to the concepts of the ‘adjacent possible’, which is used
in ecosystem theory, and the ‘successor’, which is used in set theory.
The concept of the ‘adjacent possible’ appears in an interview with Kauffman in
which he considers the existence of general laws in ecosystems. In the interview
Kauffman states:
164
Chapter 7
There is a chance that there are general laws. I've thought about four of them.
One of them says that autonomous agents have to live the most complex game
that they can. The second has to do with the construction of ecosystems. The
third has to do with Per Bak's self-organised criticality in ecosystems. And the
fourth concerns the idea of the adjacent possible. It just may be the case that
biospheres on average keep expanding into the adjacent possible. By doing so
they increase the diversity of what can happen next. It may be that biospheres,
as a secular trend, maximise the rate of exploration of the adjacent possible. If
they did it too fast, they would destroy their own internal organisation, so there
may be internal gating mechanisms. This is why I call this an average secular
trend, since they explore the adjacent possible as fast as they can get away with
it.
The concept of the successor can be found in mathematics, where the
successor of the number α is defined as the smallest number greater than α.
When considering numbers, this is a sharp definition. In principle, we would like
to obtain the same accuracy when translating the successor definition to the field
of particle hierarchy and closure, for example, in the form of the following
tentative definition: the successor of a particle A is the particle showing the least
complex closure higher than A. However, before one can regard the ‘least
complex closure higher than A’ as the ‘successor closure’, one must find a way
to indicate the level of closure that is simultaneously simple enough and
complex enough to be considered the successor closure.
Closure
General mathematical aspects of closure were already discussed in chapter one
where it was also explained that the closure concept in the operator theory is
used more specifically than it is in mathematical theory. In the operator theory,
closure relates to a state of matter that results from a self-organisation process
and that shows a closed topology for structure, called structural closure, or
process, called functional closure, or the combination of both, called structurofunctional closure. When used in this way, closure always represents a finite set
of relations, but as long as structural closure is absent, the relations may not
define a physical system limit. Furthermore, the state and the process of closure
are intimately related because a closure state is caused by underlying
interactions.
Elementary closure
In real systems, closure may be embedded in complex interactions. If all details
of these interactions were considered, it would become too complex to compare
closures between systems. For this reason, the operator theory uses a
methodology that reduces such complexity by identifying the minimum
Towards consistent definitions
165
complexity that still represents the system’s closure type. The latter minimalistic
representation is regarded as the elementary closure.
An elementary closure represents the lowest complexity realisation of its type for
the system’s structure, process, or when possible, the combination of both. The
elaboration or repetition of the same closure type does not affect the elementary
closure. For example, the closure by which a chain of carbon atoms creates a
cyclic molecule (e.g., benzene, fullerene) is not regarded as creating a new
elementary closure because the closure repeats the covalent binding of atoms;
the covalent binding was already responsible for the self-organisation process
that produced molecules from atoms. Accordingly, every molecule is a molecule,
regardless of its construction. Similarly, the adding of electron shells to an atom
and the presence of more than one nucleus in a multinucleate cell does not
change the elementary closure. Finally, elaborating a catalytic cycle with shunts
and connections between shunts will not affect the elementary closure type of
catalytic closure.
A system loses its elementary closure when the construction and/or dynamics
are reduced beyond the minimum situation required for a closed state, for
example, an atom that is heated until it loses the last electron from its electron
shells or an organism starving to the level where it cannot maintain autocatalysis
in its minimum form and dies. To sustain elementary closure, a system must be
capable of a minimum level of maintenance. In the context of chemical
reactions, Dittrich and Speroni di Fenizio (2007) have specifically focused on
maintenance by defining an ‘organisation’ as any ‘set of molecules that is both
closed and self-maintaining’.
First-next possible closure
First-next possible closure is defined as follows: given any system A that shows
first-next possible closure, the next first-next possible closure creates the least
complex system type above A that shows a new type of elementary closure
based on A and, when required, any highest level system type possible below A
that shows first-next possible closure. This definition is inherently recursive
because a system showing first-next possible closure produces a more complex
system showing first-next possible closure. Because a first-next possible closure
is always built from systems showing a preceding first-next possible closure, the
recursive definition does not lead to logical loops.
The above definition implies that any system showing first-next possible closure
is produced by a system showing one lower level of first-next possible closure.
Going down this ‘staircase’ either results in an infinite regression or suggests the
existence of a least complex system that shows first-next possible closure. A
regression ad infinitum represents an undesired perspective for real world
systems because every step down the closure-staircase is accompanied by a
166
Chapter 7
loss of organisational complexity and a reduction of the possibilities for closure.
This regression strongly suggests that a least complexity system exists in nature
that shows first-next possible closure without showing a substructure of lower
level closed elements.
In general, the first-next possible closure of the system above A will be based on
elements showing the same type of first-next possible closure as A. In certain
cases, however, the construction of a higher level may also involve elements
residing at more than one level below the newly formed element. For example,
the atom is based on interactions between hadrons creating the nucleus, while
the electrons of the electron shell show a lower closure level than hadrons. As
the electrons are the highest complexity solution for creating closure nature
allows, this situation still represents a valid first-next possible closure.
Exhaustive closure
The definition of first-next possible closure makes use a new type of elementary
closure with respect to a hypercyclic process and/or its interfacing. To identify
the new type of elementary closure, I suggest focusing on ‘exhaustive closure’ in
the system at the one lower closure level. A system shows exhaustive closure, if
there is no remaining potential for elaborating the structural and dynamic
aspects of the elementary closure type, because any further development will
cause the construction of a new elementary closure type. While the elementary
closure of the system remains the same, the state of exhaustive closure allows
for a new functionality.
Exhaustive closure can be applied at all the levels in the operator hierarchy. For
example the transition to hadrons makes use of the relatively complex quarks
and the transition from the multi-atoms to the autocatalytic sets makes use of
three dimensional catalytic potential of complex molecules. With respect to the
internal differentiation that forms the basis for the transition to the
endosymbiontic/eukaryotic cell, exhaustive closure suggests that a cell should
possess enough complexity to allow for the engulfment of an endosymbiont
and/or the construction of internal membranes.
Exhaustive closure furthermore offers a guiding principle when defining the
transition from a unicellular to a multicellular state. During this transition there
are several not exhaustive configurations that could be classified as a next
closure state. Firstly, two cells can mutually attach to each other by means of
binding proteins. Secondly, the cells can develop a mutual dependency based
on the diffusion of chemicals across their membranes. Thirdly, the membrane
separating the cells could become perforated, connecting the plasma of the
cells. Further unification would create a single cell with two nuclei. Of these four
examples, the third option, where the soma of both cells is connected by pores,
represents the most exhaustive closure situation. According to the demand of
Towards consistent definitions
167
exhaustive closure, the third option represents the preferred choice; the first two
options are not exhaustive, and the fourth option implies a multinucleate cell,
which repeats the closure represented by the nucleus. Based on exhaustive
closure a strand of attached algae that do not show pores connecting the
plasma, must be considered a multicellular colony, because they lack the
prerequisites for the exhaustive multicellular closure state. The latter also implies
that slight changes are required in the definition of life and of the organism as
were given in the paper of Jagers op Akkerhuis (2010a). The reason is that the
use of exhaustive closure demands that only cells connected by plasma strands
represent multicellular operators.
System type of systems showing first-next possible closure
Two systems created by the same type of first-next possible closure are
considered to be of the same system type. This definition creates subsets of the
set of all systems that was defined above.
Closure dimension
The structural and/or functional properties of a system showing first-next
possible closure can be grouped in relation to higher level types, named ‘closure
dimensions’. A system that results from first-next possible closure either shows a
new closure dimension or shows an existing type of closure dimension. The
assignment of closure dimensions depends strictly on the most recently realised
first-next possible closure of the system and not on any other property. The
following closure dimensions can be recognised:
The interface dimension
The interface dimension involves a first-next possible closure that causes a
spatially closed topology acting as an interface separating the system’s internal
dynamics from the world. For elementary particles, the interface is assumed to
be the only dimension that determines their structure. For the hadron, the atom,
the cell and the memon, the interface dimension always represents an interface
mediating the second-order cyclic processes that are associated with the
hypercycle dimension.
The hypercycle dimension
In the operator theory, the word hypercycle is used in a general way. It refers not
only to the catalytic hypercycle but also to other second–order cyclic processes
resulting from first-next possible closures. The first time that a first-next possible
closure results in a system showing the hypercycle dimension is when quarks
and gluons form a second-order interaction cycle in quark-gluon soup. Other
dynamics that show the interface dimension are the quark confinement, the
electron shell, the cell membrane and the sensory interface of the memon.
168
Chapter 7
The multi-state dimension
The multi-state depends on a first-next possible closure that creates recurrent
interactions between systems sharing the same (one level lower) first-next
possible closure. The multi-state dimension is introduced by the hadrons, where
it depends on interactions between quarks. Other systems that are based on
first-next possible closure and that show the multi-state are the hadron, the
molecule, and the prokaryote and eukaryote multicellular.
The hypercycle mediating interface dimension (HMI)
This dimension marks the moment at which an interface is formed surrounding
and mediating the interactions of a second-order interaction process (the
hypercycle), while both interface and hypercycle result from first-next possible
closure.
The structural (auto-)copying of information dimension (SCI)
This property states hat a system’s dynamics create physical copies of the
elements responsible for the system’s emergent hypercyclic process. The first
system that shows this property as the result of first-next possible closure is the
cell, where the molecules that allow for autocatalysis represent the cell’s
information that is autonomously copied by the cell’s chemical reactions. So far,
the cell is the only system with this closure dimension.
The dimension of structural auto-evolution (SAE)
This closure dimension is associated with systems that show the capacity to
autonomously modify the structure of their hypercyclic interactions. This property
is selectively present in organisms with a hypercyclic neural system with
interface. Currently, SAE typically relates to animals.
Primary system
A system exhibiting/introducing a new closure dimension is regarded as a
primary system. Accordingly, only the following systems showing first-next
closure are primary systems: the elementary particles, the second-order quarkgluon cycles, the hadrons, the atoms, the prokaryotic cells and the memons.
Major transition
Any first-next possible closure creating a primary system is regarded as a major
transition.
Minor transition
Any first-next possible closure creating a system, the closure type of which does
correspond to an existing closure dimension, is considered a minor transition.
The sequence connecting first-next possible closures may split into different
branches when a system allows for minor transitions belonging to different
Towards consistent definitions
169
closure dimensions. An example of this is the transition from bacterial systems
to either eukaryotic cells or prokaryotic multicellular life-forms. The eukaryotic
cells show an internal closure representing the hypercycle mediating interface
dimension. The prokaryotic multicellulars show an interactive closure belonging
to the multi-state dimension. Another example is the step from molecules to
either autocatalytic sets or cellular membranes.
Complexity of a closure dimension
The complexity of a closure dimension is based on the number of major
transitions that were required to construct it. Because closure dimensions occur
in a strict sequence, their complexity level can also be numbered from 1 to 6,
from the interface dimension to the structural auto-evolution dimension,
respectively.
Pre-operator interaction systems
Systems showing closure dimension 1 or 2 are called pre-operator interaction
systems.
Operator
An operator is defined as a system type showing a closure dimension of 3 or
higher (Fig. 7.1). From a different perspective, operators can also be defined as
those systems showing first-next possible closure that are not pre-operator
interaction systems. The set of the operators unifies all particles whether they
are physical particles or organisms. The following operators are currently known:
the hadron, the atom, the molecule, the prokaryote cell, the eukaryote cell, the
prokaryote and eukaryote multicellular and the neural network organism, called
the ‘memon’ in the operator theory.
Operator hierarchy
The operator hierarchy is a conceptual ranking of all operators and pre-operator
interaction systems. The structure of the operator hierarchy is based on an
integration of the viewpoints of first-next possible closure and closure
dimensions (Fig. 7.1). When discussing the operator hierarchy, I generally
include the quarks, even though these are strictly no operators but pre-operator
interaction systems.
170
Chapter 7
Figure 7.1. The figure shows the evolution of particles of increasing complexity as
represented by the operator hierarchy. The black line shows the historical pathway by
which first-next possible closures create the operators. The gray columns indicate preoperator hypercyclic sets and interfaces. Explanation of abbreviations: SAE ('Structural
Auto Evolution’) = the property of an operator to autonomously evolve the structure that
carries its information. SCI (‘Structural Copying of Information’) = the property of an
operator to autonomously copy its information (genes, learned knowledge) by simply
copying part of its structure. HMI (Hypercycle Mediating Interface) = a closure creates an
interface that mediates the functioning of the hypercycle. Multi-state = operator showing
closure between multiple units of exactly one lower closure level. Hypercycle = closure
based on emergent, second-order recurrent interactions. Interface = closure creating an
emergent limit to an operator. CALM (Categorizing And Learning Module) = a minimum
neural memory.
Hierarchical layer
Operators and pre-operator interaction systems that are based on the same
primary system and do not yet represent the next major transition are
considered belonging to the same hierarchical layer. This means that the
operator theory recognises as many hierarchical layers as there are primary
systems.
Towards consistent definitions
171
Interaction system
An interaction system consists of operators that interact without that the
interactions create first-next possible closure. The presence of first-next possible
closure would make the interacting elements an operator.
Operator hypothesis
The operator hypothesis states that the fundamental limitations imposed by firstnext possible closure act as a strict mould for particle evolution.
Operator theory
The operator theory is the theoretical framework that offers the context for the
operator hierarchy and the operator hypothesis.
172
Chapter 7
Chapter
8
Centripetal science
Contributions of the operator hierarchy to
scientific integration
“Meta-plumbing would concentrate on such questions as
how to look as though you really understand what you are doing
when the living room is several feet deep in water
- instead of how to make a good joint, which is ordinary plumbing.”
(Jack Cohen and Ian Stewart)
174
Chapter 8
Summary
The large number of discoveries in the last decades has caused a scientific
crisis that is characterised by overspecialisation and compartmentalisation. To
deal with this crisis, scientists need comprehensive approaches and unifying
theories. A recent integrating theory, the operator hierarchy, connects the
evolution of physical particles with that of organisms. In this chapter it is
investigated what novelty the operator theory can add to a range of already
existing integrating approaches. The following approaches will be discussed: the
use of a cosmic timeline, the ranking of systems according to an ‘ecological
hierarchy’, the use of periodic tables, the unification of evolutionary processes by
regarding evolution as an artesian well, the generalisation of the evolution
concept and a cross table plotting the applicability of unifying concepts against
the different complexity levels that the operator theory recognises for ranking
system types.
Centripetal science
175
Introduction
The flow of scientific information is increasing daily and is causing what could be
considered a knowledge explosion crisis (Sagasti, 1999). Dealing with this crisis
requires not only methods that handle and make accessible huge amounts of
information but also integrative theory that offers scaffolds to connect distant
scientific fields. This text focuses on the role of the operator hierarchy as a
scientific integration tool in the context of other integrative principles.
The knowledge explosion crisis is the result of science having a positive
feedback on its own development. Science depends on our capacities as
humans to observe the world, including ourselves, and to create mental
representations of the observed phenomena. In turn, better representations
increase our capacities to manipulate the world and to construct tools that
improve our observations, which, in turn, accelerate scientific development. As
the result of this process, scientific ideas have simultaneously developed
extreme breadth and depth. They are not only frequently associated with
overspecialisation and compartmentalisation but they are also moving towards
the far reaches of major scientific integrations of a cosmic scale. At any level
between these extremes, a concept is more valuable when it is efficient, pairing
minimum complexity with maximum precision and elegance. Arbitrage by these
aspects has become known as “Ockham’s razor”, cutting away the most
complex and least elegant from any two theories explaining the same
phenomenon. If theories describe different phenomena, arbitrage is less
straightforward because different viewpoints may highlight different aspects of
natural organisation. Considering the latter points, this paper analyses how the
operator hierarchy may contribute to scientific integration while focusing on
major integrative ideas, for example, the use of timelines, ecological hierarchy
and evolution. The paper ends with an overview of existing integrating theories
and the value of scientific integration.
Unification based on complexity and timelines
In history various theories have been developed to create coherence in the
phenomena observed in the world. The structure of such theories has focused
on the complexity of systems or has centered on timelines for historical ranking.
A classical example of a linear hierarchy ranking system complexity is the Scala
Naturae of the Greek philosopher and naturalist Aristotle. In his approach,
Aristotle ranks natural phenomena by decreasing perfection, from spiritual and
divine beings to man, animals, plants and finally rocks and formless matter. This
classification is also referred to as the Natural Ladder or the Great Chain of
176
Chapter 8
Being and has influenced medieval Christianity and the thinking of well- known
philosophers such as Descartes, Spinoza and Leibniz.
An approach that dates even further back in history is the ancient symbol of the
Ouroboros, the snake swallowing its tail. This symbol represents the continuity
and eternity of life and the world. A recent interpretation of the Ouroboros is
shown in Figure 8.1. The left half of this figure shows how science has gained
access to increasingly small systems, from large animals and plants, to
individual cells, to molecules, to atoms, to nuclear particles and finally to
fundamental particles. The figure’s right half shows the scope of scientific
research broadening towards increasingly large systems, from ecosystems,
through planets, to solar systems, galaxies and finally the entire universe. When
presented in this way, specific properties of the universe at large, such as the
background radiation and its expansion in all directions, are connected with the
universe’s initial state. As the picture of the Ouroboros suggests, when the very
large and the very small meet, the search for organisation in nature seems to
swallow its tail.
Figure 8.1. Historical expansion of scientific domains in the direction of the smallest
systems and in the direction of the largest systems. GUT = grand unifying theory, DM =
dark matter, W and Z indicate the particles conveying the weak nuclear force, cm =
centimetre.
In addition to ranking systems by complexity, systems can also be ranked by the
moment of their first formation and the historical time period in which they
existed. When analysing system organisation in this way, timelines at different
scales are created that refer to, for example, palaeontology, particle physics,
Centripetal science
177
human history, the development of the automobile. These timelines also come in
different forms, such as linear hierarchies, branching trees and circular
approaches.
A modern timeline presenting a comprehensive overview of the organisation in
nature is Big History (Spier, 1996; Chaisson, 2001; Bryson, 2003; Stokes Brown,
2007). This approach ranks all systems and processes by their occurrence in
cosmic history. Big History is based on the scientific theory that the early
universe was as small as a fundamental particle and obtained its present size
following a rapid expansion, the Big Bang. The theory that the universe has a
minute origin is supported by modern particle physics and by cosmological
observations of the background radiation and the proportionality with distance of
the speed at which galaxies recede from the Earth in all directions (Pagels,
1985; Weinberg, 1977). Based on these observations, it has been calculated
that Big History started about 13.7 billion years ago (Fig. 8.2). During the
universe’s first three minutes, quarks formed and then condensed to form
hadrons (such as protons and neutrons). During the following 17 minutes, the
hadrons condensed to form simple helium nuclei (the combination of a proton
and a neutron). After these initial minutes, it took about 70 thousand years
before the dynamic balance of the transformation of matter and energy toppled
to the advantage of matter. The matter in the rapidly expanding universe now
aggregated under the influence of gravity. The aggregation process was slow
because gravity is weak at large distances. The result was a universe with a
sponge-like structure of concentrations of matter surrounding empty ‘bubbles’ of
variable size that were almost devoid of matter. After 100 million years of
aggregation, the first galaxies and stars were formed and their light started
illuminating the universe. The nuclear reactions in stars supported the formation
of elements heavier than helium. After approximately 9.1 billion years, the Sun
was formed (4.57 years ago) and then planet Earth (4.54 billion years ago).
Thereafter, it took about 1 billion years for the first life to emerge on Earth and
another billion before cells gained the capacity of photosynthesis. Complex
Ediacaran fauna has been found in rocks of about 600 million years old. Around
228 million years ago dinosaurs ruled the world. The first hominin fossils
originate from approximately 7 million years ago. Human history dates back to
several tens of thousands of years.
A universal timeline is a comprehensive integration tool. Its major strength is
ranking all sorts of events simultaneously. Even though every event has only a
single moment of occurrence, a timeline can flexibly adapt to variations in the
moment of occurrence of similar events by indicating a first moment at which
they occur and, assuming that such is known, a last moment. A universal
timeline can thus be seen as a thickly woven cable of many threads representing
local histories and developmental rates of different parts of the universe. (Fig.
8.2). Although all of these developmental threads unroll in different directions,
178
Chapter 8
they result in similar histories. Stars are formed everywhere in the observable
universe and their formation roughly started at the same moment. Stars of the
same class also consist of similar particles and atoms, and stars the size of the
sun are probably circled by planets everywhere in the universe. One may now
ask why there is so much uniformity in the universe and whether such uniformity
may be used to answer questions about the future of the universe.
Figure 8.2. Universal timeline (modified from Wikipedia).
A methodology that may offer answers to the latter question is the operator
hierarchy (Jagers op Akkerhuis and van Straalen, 1999, Jagers op Akkerhuis
2001, 2008). The operator hierarchy combines system complexity with a timeline
of system formation. To explain this approach, it is useful to start with the basic
Centripetal science
179
assumption that nature must be analysed according to three fundamental
dimensions for structural complexity. As Figure 8.3 illustrates and taking a
prokaryotic cell as a starting point, the first dimension involves the ranking from
a single cell to a system with interacting cells. Systems like this, which consist of
interacting organisms, were named ‘interaction systems’ by Jagers op Akkerhuis
and van Straalen (1999). The second dimension involves the transition from the
prokaryote cell to the eukaryote cell and then to the multicellular. By extending
the ranking in this dimension towards systems of low complexity, such as
molecules, atoms, hadrons and fundamental particles, it becomes clear that this
dimension connects all physical particles and organism types. Because all
elements in this dimension are regarded as ‘operators’ (Jagers op Akkerhuis and
van Straalen, 1999; Jagers op Akkerhuis, 2008), their hierarchical ranking is
called the operator hierarchy and the related theory the operator theory. As
Figure 8.3 shows, transitions towards higher level operators may use
interactions between lower level operators (e.g., from unicellular to multicellular),
but they may also involve internal differentiations (e.g., the engulfment of an
endosymbiont in a eukaryotic cell and the formation of a neural network in a
multicellular). The third dimension involves the cell’s/organism’s interior
organisation. Here the focus is on the elements, the elements in these elements,
etc. The above dimensions capture independent directions for analysing natural
organisation: a prokaryotic cell, a eukaryotic cell and a multicellular organism
can all be involved in ecosystem interactions and each of them shows internal
organisation.
Figure 8.3. Three independent dimensions for hierarchy in the organisation of nature: (1)
hierarchy in interaction systems (systems that consist of operators without being an
operator), (2) hierarchy in the way how lower level operators create higher level
operators, and (3) hierarchy in the internal organisation of operators. Only the hierarchical
ranking of the operators is strict. All other hierarchies vary according to point of view, e.g.,
displacement, information, construction and energy.
180
Chapter 8
Having explained how the operator theory recognises dimensions for
organisation, it is time to return to the question of how these insights can assist
in analysing cosmic development. The answer to this question rests on
separating the events along the cosmic timeline into two parallel tracks: the track
of the operators and the track of the interaction systems.
Starting with the Big Bang, the history of the universe can, in principle, be
modelled as a container full of interacting particles. Particles exert forces on
each other and interact. This interaction forms new particles and accompanying
new forces. During the universe’s early days, the initial quark soup rapidly
transformed to also contain simple helium nuclei. Simultaneously, and on a
larger scale, all matter aggregated due to the force of gravity. This aggregation
created various celestial bodies, such as black holes and stars. Nuclear
reactions in stars then fused helium to heavier elements, which were spread by
stellar explosions. Under colder conditions the fused elements further
condensed to form molecules that formed a basis for planets and comets.
Models predicting the future of this process have been based on the total
amount of matter, the gravitational constant, the expansion rate of the universe,
and the life histories of celestial bodies. Although uncertainties exist about the
values of certain parameters, such models generally predict the universe’s heat
death as the consequence of diluting matter in the vastness of an extremely
large, cold space.
In comparison to the latter universal models, the operator theory (Jagers op
Akkerhuis and van Straalen, 1999; Jagers op Akkerhuis, 2001, 2008) may seem
to focus on minor details when it introduces a strict ranking of all operators, from
quarks, through hadrons, to atoms and molecules, prokaryote cells, eukaryote
cells, prokaryote and eukaryote multicellular organisms and neural network
organisms (referred to as ‘memons’) (Fig. 8.4). Nevertheless, the first-next
possible closure seems to strictly limit the structures of the operators such that
there are important consequences for natural organisation, even at a universal
scale.
Firstly, the operator hypothesis suggests that the limits set by first-next possible
closure apply to all operators anywhere in the universe. Accordingly, the same
classes of operators can be expected to exist anywhere in the universe as long
as local conditions allow for their formation. In addition to uniform initial
conditions in the universe during its early hours, first-next possible closure rules
offer an explanation for the uniformity of the structural developments in
unconnected local parts of the universe.
Secondly, the sequence of first-next possible closures and related operators is
not only strict, but it also shows an internal regularity (Fig. 8.4). Based on this
regularity, the operator theory now suggests that it is possible to extrapolate the
Centripetal science
181
operator hierarchy towards future operators, the next of which is the memon (the
generalised concept for operators with neural networks) that owes its
intelligence to a programmed neural network (see Jagers op Akkerhuis, 2001).
Because cosmology focuses on the universe at large, cosmological models offer
no possibilities for predicting future operators.
Figure 8.4. This figure illustrates the evolution of the operators. The black line shows the
historical pathway of subsequent first-next possible closures and related operators. The
gray columns indicate systems resulting from first-next possible closure but are not
operators. Explanation of abbreviations: Memon = operator showing a hypercyclic neural
network with interface, SAE ('Structural Auto Evolution’) = the property of an operator to
autonomously evolve the structure that carries its information, SCI (‘Structural Copying of
Information’) = the property of an operator to autonomously copy its information (genes,
learned knowledge) by simply copying part of its structure, HMI (Hypercycle Mediating
interface) = a closure creates an interface that mediates the functioning of the hypercycle,
Multi-state = operator showing closure between multiple units of exactly one lower
closure level, Hypercycle = closure based on emergent, second order recurrent
interactions. Interface = closure creating an emergent limit to an operator, CALM
(Categorizing And Learning Module) = a minimum neural memory.
182
Chapter 8
Unification based on ecological hierarchy
An integration tool widely used in ecology, ecotoxicology and biology is
ecological hierarchy. The logic behind this integration tool is that elements at
lower levels in the hierarchy contribute to the systems at higher levels in the
hierarchy. The general pattern of the ecological hierarchy is shown in Figure 8.5.
With slight variations, comparable ecological hierarchies can be found in many
basic textbooks and publications on ecological organisation (e.g., Odum 1959;
Weiss, 1971; Koestler, 1978; Close, 1983; Kruijf, 1991; Haber, 1994; Naveh &
Lieberman, 1994; Hogh Jensen, 1998; Korn, 2002). The ecological hierarchy’s
frequent occurrence in publications shows that this integration tool has worked
so well that it has become a kind of dogma in ecology and biology. As a
consequence, people seem to accept it unquestioningly. In the context of this
paper, the ecological hierarchy is analysed in comparison with the three
hierarchical dimensions of the operator theory explained above.
Figure 8.5: The classical ecological hierarchy
The application of the operator theory now dramatically impacts the ranking of
the elements of the conventional ecological hierarchy (Fig. 8.6). What has
happened?
Centripetal science
183
Figure 8.6. Ecological hierarchy following the ranking in three dimensions that is
suggested by the operator theory
The most important cause of the observed differences is that the operator theory
does not accept mixing hierarchal dimensions. This means that it does not allow
switching between hierarchy rules when going from one level in the hierarchy to
the next. For example, after the hierarchy of operators has been selected for
analysis, e.g., atom-molecule-prokaryote cell, an organ or a community of cells
can no longer be included as the next level. Including an organ would confound
the operator dimension with the internal complexity dimension. Including a
community of prokaryotic cells would confound the operator dimension with the
dimension that is reserved for hierarchy in interaction systems. Interestingly, the
conventional ecological hierarchy ends with the universe as the highest level,
while the operator theory recognises the universe as the first occurring
interaction system.
With the passing of time, and in close relationship with the formation of higher
level operators, local parts of the universe change organisation and become
identifiable as subsets with specific properties. The first subsystems that develop
in the universe are concentrations of matter, parts of which further develop into
primordial galaxies. Within such galactic concentrations of matter, gravity causes
the atom nuclei that have formed to aggregate locally and form primordial solar
systems. As a next step stars and planets may form. As soon as cells are
formed on a planet, this planet changes from a chemosystem to an ecosystem.
Subsequently, when memons arise on a planet, this development can be
regarded as changing the planet into a ‘memosystem’.
184
Chapter 8
Finally, the operator hierarchy demands that the hierarchy of elements in an
operator’s interior be dealt with independently. In abiotic operators, such as
atom nuclei, atoms and molecules, the internal differentiation directly results
from interactions based on condensation (from hadrons to nuclei, from nuclei
and electrons to atoms, and from atoms to molecules). For organic operators
Turchin (1977) has formulated the law of the branching growth of the
penultimate level. This law states that '...after the formation, through variation
and selection, of a control system C, controlling a number of subsystems Si, the
Si will tend to multiply and differentiate’. This law explicitly recognises that only
after the formation of a mechanism controlling the subsystems Si is there a
context that allows the variety of the subsystems to increase. Accordingly,
nature has had neither a context nor the means to develop organelles before
cells or to develop organs and tissues before multicellular organisms. For this
reason, including sub-systems such as organelles, tissues, organs and organ
systems in the conventional ecological hierarchy of systems is highly confusing.
Unification based on a periodic table of periodic tables
The most well-known periodic table is the periodic table of the elements.
Mendeleev introduced this tabular display of the chemical elements in 1869. It
organised the reactivity of atoms and indicated a number of missing elements.
Mendeleev’s discovery was so important that his table is still used as a basic
tool in chemistry.
But chemistry is not unique when it comes to periodic tables. Various periodic
tables of fundamental importance exist for other disciplines. Probably the most
well-known is the ‘standard model’ used in particle physics. It categorizes the
major classes of fundamental particles as either force carrying particles (bosons)
or matter particles (fermions). The fermions are subsequently divided into
leptons or quarks, both of which are partitioned over three groups of increasing
mass.
Another fundamental periodic table used in particle physics is the ‘eightfold way’.
This table is used to organise the many ways by which quarks can combine into
hadrons. Hadrons consisting of two quarks are called mesons while those made
up of three quarks are called baryons, and a separate table exists for each of
these types. The eightfold way was developed by Gell-Mann and Nishijima and
received important contributions from Ne'eman and Zweig (Gell-Man and
Neeman 1964).
Furthermore, two tables can be considered the fundaments of Mendeleev’s
periodic table: the ‘nucleotide chart’ and the charts showing which sets of
electron shells are to be expected in relation to a given number of protons.
Centripetal science
185
Finally, and even though it may be a bit unusual to regard this arrangement as a
periodic table, there are also good grounds to include the ‘‘tree of life” in this
overview of tabular presentations. The only difference with the other tables is
that the tree of life also includes descent, a property that has no meaning in the
other periodic tables discussed so far. In all other aspects, the tree of life has
similar properties of creating a unique and meaningful overview of all basal
types of operators, which enter the scheme as species.
Every single periodic table discussed above is central to its proper field of
science. But the tables are not connected. The operator theory, however, shows
that it is possible to connect the separate tables by focusing on the types of
elements in every table. If this is done, the operator hierarchy can be used as a
‘periodic table for periodic tables’ to organise the elements of the existing
periodic tables.
As Table 8.1 shows, the inventory of periodic tables offered a periodic table for
almost every complexity level in the operator hierarchy. The inventory
furthermore indicated the following gaps for which no periodic tables were found:
the quark-gluon hypercycles, the quark confinement, the molecules, the
autocatalytic sets, the cellular membranes, the cyclic CALM networks and the
sensory interfaces. With the exception of the molecules, which may not have a
periodic table because of the almost unlimited number of combinations that can
be made from the various atom species, all the gaps involve hypercyclic sets
and interfaces. One may now suggest that it is generally impossible to create
periodic tables for hypercyclic sets or for interfaces, but this assumption is at
least partially contradicted by the nucleotide chart and the classification of
potential electron shells. A reason for the absence of tables for hypercyclic sets
may be that the number of possible configurations is so large that it is impossible
to classify them, in the same way that it is hard to classify molecular
configurations. Such ideas, however, need to be worked out in more detail.
186
Chapter 8
Table 8.1. Using the operator hierarchy for organising the periodic tables that exist for
different types of operators. Shading indicates system types that are operators.
System type in operator hierarchy
fundamental particles
quark-gluon hypercycle
quark confinement
hadrons
nuclear hypercycle
electron shell
atoms
molecules
autocatalytic sets
cellular membrane
cells
eukaryote cells
prokaryote multicellulars
eukaryote multicellulars
cyclic CALM networks
sensors (perceptive and activating)
memons (hardwired)
memons (softwired)
Organisation in specific ‘periodic table’
standard model
??
??
eightfold way
nucleotide chart
types of shells
periodic table of the elements
??
??
??
tree of life: prokaryotes
tree of life: eukaryotes
tree of life: prokaryote multicellulars
tree of life: eukaryote multicellulars
??
??
tree of life and technical hardwired memons
future tree of technical life
Unification based on organic evolution: the artesian well that
is powered by cellular autocatalysis
Calvin (1987) describes evolution as a ‘river that flows uphill’. Dawkins (1996)
refers to it as a process that is ‘climbing mount improbable’. Neither of these
metaphors suggests the need for a force to realise the process. To clearly
indicate that a driving force is needed to make water flow against gravity or to
make evolution climb a mountain, the metaphor of an artesian well will be used.
In an artesian well, the groundwater pressure makes the water flow naturally
towards the surface allowing it to ‘defy’ gravity. But what exactly is the pressure
that makes evolution flow towards increasing complexity, seemingly ‘against’
thermodynamic laws? As Russell (1960) and Pross (2003) have indicated, this
pressure is a special form of the explosive, brutal power of autocatalysis. Taking
Pross’s insight as a basis, the following text places evolution in a thermodynamic
perspective and invokes the operator hierarchy when appropriate.
Long ago Malthus (1798) and Verhulst (1838) realised that population growth
leads to density dependent stresses. Darwin (1859) subsequently developed the
idea that this stress, in combination with reproduction and heritability of parental
properties, causes reproductive disadvantage of the least adapted individuals.
Centripetal science
187
However, Darwin and contemporaries had no clear idea about what could cause
the organisation of organisms. The laws of thermodynamics that were known at
that time seemingly indicated that systems could not increase their organisation
(Carnot, 1824; Clausius, 1865). A bit later, Bergson (1911) wrote about life:
Incapable of stopping the course of material changes downward (the second law
of thermodynamics), it succeeds in retarding it …Now what do these explosions
(photosynthetic reactions) represent, if not a storing up of the solar energy, the
degradation of which energy is thus provisionally suspended on some of the
points (the plants) where it was being poured forth?
Later, ideas about non-equilibrium thermodynamics (Schrödinger, 1944;
Prigogine, 1984) and hypercyclic catalysis (Eigen, 1979) offered the ingredients
for a better explanation. Non-equilibrium thermodynamics solved the problem
that growth and reproduction seemed to violate the laws of thermodynamics.
What was new in open thermodynamic systems was the idea that the
degradation of an external free energy gradient could power the dynamics
required for self-organisation. For example, when a bathtub is unplugged, the
self-organisation of the vortex is powered by the degradation of the potential
energy stored in the height difference between the water in the tub and the drain
at the bottom. But although non-equilibrium thermodynamics offered a general
solution for the powering of self-organisation, it did not indicate what specific
driving force powered evolution.
To analyse the processes that drive Darwinian evolution in more detail, evolution
will be analysed as the combination of two processes: one process explaining
the functioning of organisms, from single cells to animals, and the other process
explaining selection. The functioning of unicellular organisms requires selforganisation and a membrane. Self-organisation is powered by transforming
external energy gradients into work. As will be discussed presently, the operator
hierarchy indicates that the organism receives the storage of heritable
information for free as long as it uses hypercyclic autocatalysis as the basis for
its energetics. The membrane is required to ensure that the information and
other processes become individualised. The mechanisms behind selection
depend on the capacity to produce offspring that receive variable heritable
information, and on selective interactions affecting the phenotypes of the
offspring differentially.
The basal self-organisation process responsible for the existence of organisms
is autocatalysis. Autocatalysis in its basic form is the process in which a certain
catalytic chemical, say A, transforms a substrate, which then leads to the
production of A. Given sufficient substrate, autocatalysis leads to the doubling of
catalyst molecules with every transformation step, from A, to 2A, 4A, etc. This
process is referred to as exponential increase. The potential power of
188
Chapter 8
exponential increase can be derived from the three dynamic states an
autocatalytic process may attain (e.g., Lifson 1987; Dittrich and Speroni di
Fenizio 2007): (1) when the influx of substrate is too low, the system decays; (2)
when the inflow of substrate is high enough to let the autocatalytic production of
catalysts equal their decay rate, the system is in (dynamic) balance; and (3)
when there is a rich influx of substrate, the positive feedback causes a chain
reaction that will let the process grow exponentially. While systems with
decaying or balanced dynamics will go unnoticed, systems with exponential
growth potentially possess the brutal force of an explosion.
The explosive power of autocatalysis is not sufficient to explain Darwinian
evolution because autocatalysis lacks heritable information. The coupling of
autocatalysis and information requires an additional step. In its simplest form this
second step requires the coupling of two catalytic reaction cycles based on the
molecules A and B in a second-order cycle in which A transforms substrate to B
and B transforms substrate to A. The resulting reaction cycle is fully driven by an
external free energy gradient and is a simplified form of Eigen’s ‘catalytic
hypercycle’ (Eigen, 1979). Eigen, who focuses on enzymatic reactions, has
published various studies about the stability and thermodynamics of hypercyclic
catalysis. In a catalytic hypercycle, every individual catalytic molecule can be
regarded as carrying information for the overall process. The capacity of
hypercycles to carry information has recently been discussed by Silvestre and
Fontanari (2008). The hypercycle thus combines the explosive force of
autocatalysis with the information function of the separate catalytic molecules.
Hypercyclic catalysis unleashes enormous powers while creating an informed
process. However, these properties are still insufficient to cause evolution
because the process does not yet include a spatial mediating boundary that
allows the components to become a unit of selection. Without a boundary, the
catalysts of an autocatalytic hypercycle float freely in the pre-biotic ‘soup’ and
cannot be assigned to a specific group. They can dilute or mix freely with other
sets. To end up with units that selection can act on, a physical system limit is
required. This can be added quite easily as a fatty acid membrane. Vesicles
naturally form by condensation in a watery solution containing fatty acids
(Oparin, 1957; Claessens et al., 2007; Fanelli, 2008; Hernández-Zapata et al.,
2009) and the process is well understood from the aspect of thermodynamics.
The combination of a membrane with autocatalysis now defines the first primitive
cell. In one of the more recent studies on the emergence of the first cells, Martin
and Russel (2003) have discussed the simultaneous formation of autocatalysis
and membranes based on the chemical reactions in pre-biotic submarine
hydrothermal vents of volcanic origin.
Once primitive cells were produced, it was a relatively small step toward
multiplication and heritability of information. Given a constant supply of substrate
Centripetal science
189
molecules, autocatalysis automatically increased concentrations of the catalytic
molecules in a cell. It also potentially produced fatty acids enlarging the
membranous envelope. Increasing cell volume and envelope size will destabilise
the cell structure and stimulate division, and the contents will then more or less
be randomly distributed over the two ‘offspring’. When this occurs, cell-based
autocatalysis powers the motor of primitive cellular reproduction and only
selective interactions have to be added before evolution occurs.
Despite their primitive state, the above cell-based reproduction and heritability
immediately force the water in the artesian evolution well to flow upward. The
reason is that cell-based cyclic autocatalysis implies the production of numerous
individuals, that the individuals show interactions and that interactions are most
detrimental for weak performers, which Darwin referred to as the ‘less well
endowed’. The latter processes result in selective interactions that scaffold the
development of increasingly complex building plans (at least on average).
Selective interactions, as used here, are not limited to competitive interactions
but also include strategies based on cooperation.
Information, first in the form of the set of autocatalytic molecules and later in the
form of RNA/DNA, plays an important role in evolution. A fundamental aspect of
information is that it is hard to avoid random changes during its use and/or
reproduction. As a consequence, the information in organisms naturally tends to
change over generations. A negative result of this uncontrolled change is that
offspring may suffer a lethal accumulation of deleterious mutations. This kind of
mortality is referred to as Muller’s ratchet (the name is derived from the random
occurrence of deleterious mutations as discussed by Hermann Joseph Muller,
1890-1967). A positive result of this change is that every once in a while a given
mutation will positively affect an organism’s fitness. As long as the production of
original types and mutants that fit equally well or better to their environment
outweighs deleterious mutations, evolution will continue. The potential for
genetic evolution has convincingly been demonstrated in experiments that
investigated how the genetic material of viruses adapted over generations when
it was subjected to different chemical stresses (Mills et al., 1967; Spiegelman,
1971).
The constant emergence and spread of favourable mutations unpredictably
changes the ecosystem. To maintain one’s fitness, units of selection must
continuously adapt. The continuous need for adaptation has been simulated by
Schneppen and Bak (Bak 1996) who in a group of competing species replaced
the least fit species by one that is more fit (Bak, 1996). Their model showed that
the resulting dynamics are inherently unpredictable. This was concluded from
the fact that when plotting the number of species involved in one extinction event
against the frequencies of such events, their model showed the fractal
characteristic of a power law distribution. Such a power law accorded well with
190
Chapter 8
the distribution of species’ extinctions in the archaeological records, as van
Valen (1973) observed. After changing the original Schneppen-Bak model to be
more realistic, for example, by including genetic adaptation and random
disturbances caused by meteorite impacts, the model was proven robust and
relevant for the evolutionary process.
During evolution, selection acts not only in the direction of the capacity to evolve
but also of the capacity to evolve evolvability (e.g., Wagner, 1996; Wagner and
Altenberg, 1996). Once evolution has started, it becomes increasingly difficult to
stop the process because selection will favour organisms that can exploit
formerly inaccessible free energy gradients. Every new pathway implies a new
kind of ‘fuel’ powering new autocatalytic processes and increasing the size of
and/or the pressure under evolution’s artesian well. Examples of switches
towards new and larger free energy gradients are those from physico-chemical
energy to solar energy (the development of photosynthesis), from physicochemical
energy
to
biochemical
energy
(the
development
of
predation/herbivory), from anaerobic pathways to the use of oxygen (yielding a
twenty-fold increase in available energy), from the exploitation of living biomass
to the use of fossil biochemical energy, etc. Other examples are (1) the switch
from depending on diffusion for energy transport to active transportation of
energy-rich substrates through the cell and (2) the symbiosis with
endosymbionts, generating energy throughout the cell.
From the above switches, the switch from physico-chemical energy to
biochemical energy has especially affected the evolutionary process because
the biomass of the early organisms suddenly became a degradable free energy
gradient. Exploiting this gradient, viruses, parasites and consumers attacked the
organisms. These attacks reduced the densities, which, in turn, increased
growth rates. The chisel of selection was sharpened when indirect competition
for abiotic resources was supplemented by organothrophic interactions.
Afterwards, selective forces showed diversification towards searching for and
digesting biotic resources and towards developing survival strategies to avoid
becoming a resource.
Unification based on a general concept for evolution
Darwin’s theory refers to evolution as a combination of two processes: (1) the
production of numerous offspring with different combinations of heritable
properties, and (2) the selecting away of individuals that are less well endowed;
that is, in comparison to nearby organisms, their competitive and/or cooperative
properties fit less well to the demands of the abiotic and biotic environment. The
focus on these processes has linked the evolutionary process to heredity. In
actuality, however, evolution requires nothing more than repeating a process
Centripetal science
191
that combines the production of variation (a diversification step) with selection in
relation to certain criteria (a selection step) (Fig. 8.7). As has been indicated by
Popper (1972, 1999) and Campbell (1960, 1990) repeated diversification and
selection steps offer a general basis for evolution of organisms. But the
implications of diversification and selection may reach further, because these
concepts are not limited to organisms.
Figure 8.7. This figure illustrates the generalisation of the evolution concept Both the
evolution of particles and the evolution of organisms can be regarded as consisting of
steps combining the production of variation (diversification) and selection.
The production of variation may be based on genes but may also be based on
abiotic particles or computer organisms. For example, when two fundamental
particles meet, they may integrate and split again, or they may exchange a third
particle, such that after the process, two new particle types are formed. And
when a technical memon copies its brain structure through computer code,
incidental or deliberate errors in the process may produce variation.
The selection process, too, is not limited to organisms. In Darwinian evolution,
selection may occur at many points, including when two organisms choose each
other as mates, when semen search for an egg cell, when an embryo develops
in a uterus, when offspring are born and have to persuade their parents to feed
them, etc. (Fig. 8.7). In particle evolution, selection depends on whether particles
recombine and whether they produce new particles that are stable.
When examining evolution in the above way, the difference between Darwinian
evolution and the evolution of particles fades and the principle of evolution
becomes visible in its most basic and general form, which is based on
diversification and selection. Of course, to apply this basic form to a given level
in the operator hierarchy, the specifics of the diversification processes and the
192
Chapter 8
selection processes that are typical for the selected operators and environment
need to be filled in.
Unification based on unifying concepts
The above paragraphs have highlighted many grand unifying concepts in
science and have shown how the operator hierarchy may contribute to these
fields. The examples that were discussed represent a limited selection of the
many larger and smaller unifying concepts that exist. To discuss all these in a
single paper would detract from the major goal of this study: to analyse
relationships between unifying concepts and to discuss the potential contribution
of the operator hierarchy.
To analyse relationships between several unifying concepts while preventing
endless elaboration, it was decided to create a cross-table at a high level of
abstraction. On one axis the table shows an inventory of unifying concepts and
on the other their relevance at different levels of the operator hierarchy. It was
also decided to construct not one, but two cross-tables: one for unifying
concepts relating to operators and one for unifying concepts relating to
interaction systems (the systems that consist of operators but are not operators).
In addition, it was decided to sort the unifying concepts a priori according to the
four dimensions of the DICE approach (Displacement, Information, Construction
and Energy, Jagers op Akkerhuis, 2008), such that unifying concepts dealing
with similar subjects were gathered into these four classes. The outcome of
these activities is shown in Tables 8.2 and 8.3. Undoubtedly, these two tables
are not complete and the a priori assignment of a given unifying principle to one
of DICE’s four classes or the a priori assignment as being most important for
operators or interaction systems may be disputed because many principles
relate to more than one subject. It is furthermore recognised that some unifying
concepts have a narrow scope, for example, the Pauli principle, while other
principles, for example, the concept of evolution, could have been split up into a
whole range of related unifying concepts, such as the selfish gene, the moving
fitness landscape, game theory, etc. In relation to the latter remark, it has been
attempted to combine a priori smaller concepts into overarching concepts.
Concepts were separated if at least one aspect was important enough to justify
this decision. Due to these considerations, Table 8.1 and Table 8.2 should be
considered explorative tools for identifying interesting trends.
Centripetal science
193
atom nuclei
and atoms
molecules
(prokaryote)
cells
eukaryote cells
multicellular
prokaryote
Relativity, matter-energy
x
Thermodynamic laws
Autocatalytic physiology, resource dominance, energetic demands
Resource allocation and trade-offs, DEB-model
Desire based activity and flows
FNP closure defining the operator types
Functional limits irt levels of resources or (abiotic) stressors (Eyring, von
Liebig, Paracelsus). Niche concept
Selforganization and selforganized criticality
Particle-wave duality
Schroedinger wave functions for electron shell
L-systems, gnomons (internal/external addition), fractals
Constructal law
Growth and development, allometrics
Life cycle during one generation (ontogeny and life stages), fitness, trade-offs
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
x
x
x
x
x
x
x
x
Construction
Information
Displacement/dynamics
Genes, genetic basis of ontology, of bodyplan, of phenotypic plasticity
(response curves)
Immune/self-other
Genetic basis of neural network: mental traits, nature (instead of nurture)
Neural based behavior, reflexes, memory, intelligence, consciousness, mental
health, psychology, phychiatry, welbeing, satisfaction and brain-bodysensors interaction
Memes
Zwitter-ion
Turing patterns inside operators
Transportation inside organisms
memon
(softwired)
hadrons
Thermodynamics
(energy)
memon
(hardwired)
Related subjects (it concerns states or dynamics within operators)
multicellular
eukaryote
Main subject
fundamental
particles
Table 8.2. This cross-table shows an inventory of unifying concepts ranked according to the operator hierarchy.
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
(x)
(x)
x
x
x
x
x
x
x
x
x
Chapter 8
Information
m e m o n (s o ftw ire d )
Construction
Thermodynamics of overall systems or flows
Grand Unification Theory (Theory of Everything)
Relativity, space-time
Electromagnetism
Physical chemistry
Gravity
Constructal law, patterning giving high access to flows (including interaction
chains, such as foodchain), pseudo-fractal constrution.
Selforganization and selforganized criticality (Pareto-Zipf-Mandelbroth,
pseudo-fractal distributions)
Alternative stable states/critical transitions
Biochemistry
Biotope
Pauli principle
m u ltic e llu la r
p ro k a ry o te
m u ltic e llu la r
e u k a ry o te
m e m o n (h a rd w ire d )
Temperature dependence of interactions
e u k a ry o te c e lls
Thermodynamics
(energy)
(p ro k a ry o te ) c e lls
Related subjects
a to m n u c le i a n d
a to m s
m o le c u le s
Main subject
h a d ro n s
Table 8.3. This cross-table shows an inventory of unifying concepts in science and the range of
interaction systems to which the unifying concept apply. The interaction systems are arranged in order of
the most complex operator in the system.
fu n d a m e n ta l p a rtic le s
194
x
x
x
x
x
x
x
(x)
(x)
x
x
(x)
x
(x)
x
(x)
x
(x)
x
x
x
(x)
x
(x)
x
x
x
x
(x)
x
(x)
x
(x)
x
(x)
x
(x)
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
Centripetal science
x
x
Group-against-group selection
Habitat
Stress ecology: targets,buffering and plasticity (phenotypic, population and
community adaptation (PICT))
Disease syndrome (viral, bacterial, multicellular causal agent)
Social behavior: nurture, altruism, communication, moral, economy, politics,
culture, science
Lamarckian inheritance
Evolution of memons (meme-to-network)
x
Displacement/dynamics Quantum tunneling
Logistic map, Mandelbrot set, Julia set (stability, periodicity, chaos,
(x)
bifurcations)
Turing patterns and waves
Spatial occurrencce of organism interactions, neutral theory of biogeography
Transportation (moving, phoresy, goods)
x
m e m o n (s o ftw ire d )
x
x
m u ltic e llu la r
p ro k a ry o te
m u ltic e llu la r
e u k a ry o te
m e m o n (h a rd w ire d )
Pauli principle
Evolution A (diversification via production, followed by selection)
Evolution B irt moving fitness landschape ((selfish-)genes, diversification,
reproduction (sex, populations), epigenetics, units of evolution, role neutral
mutations, genotype-phenotype, defectors, game theory)
e u k a r y o te c e lls
h a d ro n s
Information
(p ro k a ry o te ) c e lls
Related subjects
a to m n u c le i a n d
a to m s
m o le c u le s
Main subject
fu n d a m e n ta l p a rtic le s
Table 8.3: Continued
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
(x)
(x)
x
(x)
x
x
x
x
(x)
(x)
(x)
x
x
x
x
x
x
x
x
x
x
x
x
x
(x)
x
x
x
x
x
x
x
x
x
x
x
x
195
196
Chapter 8
The inventory of unifying concepts in Tables 8.2 and Table 8.3 suggests two
major trends. The first trend is that only a few concepts apply to many different
levels of organisation and, in this sense, are truly unifying. One explanation for
this lies in the fact that most theories mainly apply to either operators or
interaction systems. Another explanation is that even within the separate lists of
Table 8.2 and of Table 8.3, few theories can be found that apply to all different
operators or to all different interaction systems.
Combining the information of both tables, the following unifying concepts are
relevant for all material systems:
 Gravity impacts all systems.
 Thermodynamic laws have to be obeyed by all processes in systems
(although minute disobediences due to chance effects are possible).
 The stability of all systems is limited to within a certain range of
environmental conditions.
 Self-organisation and the constructal law connect thermodynamics with the
formation of structural patterns.
 The concept of first-next possible closure allows the recognition of the
operators and the operator hierarchy, hereby offering a “periodic table for
periodic tables”.
 If systems show dynamics, these may show various patterns, such as
alternative stable states, fractal behaviour or shifts between stable, periodic
and chaotic behaviour.
The second trend that is suggested by the inventory is the divergence between
non-dissipative or dissipative systems. Of course, there is a good reason for this:
a dissipative organisation of an operator is intimately linked to properties that
allow it to organise itself while using an external energy gradient. Examples of
such properties are autocatalysis, a membrane, heritable information, growth,
demand for food or energy, etc. More detailed subdivisions can be recognised
for all levels of the operator hierarchy because every new emergent property of
an operator introduces new properties. For example, memic closure introduces
reflexes, learning and behaviour based on mental representations.
Centripetal science
197
Discussion
The operator hierarchy contributes in different ways to scientific integration.
Firstly, it allows the structuring of various fields of scientific theory by invoking
the strict ranking of the operator hierarchy. Secondly, it establishes connections
between other unifying concepts. The structuring capacity of the operator
hierarchy results from using first-next possible closure, which allows a strict
ranking of the operators. A strict ranking means that an operator cannot be
included or excluded without disturbing the entire logic of the operator hierarchy.
If a theory possesses such strictness, this can be regarded as a special kind of
beauty. For example, Einstein said the following about his general theory of
relativity that offered a strict framework for dealing with gravity, space-time and
matter-energy: “The chief attraction of this theory lies in its logical completeness.
If a single one of the conclusions drawn from it proves wrong, it must be given
up; to modify it without destroying the whole structure seems to be impossible
(from Weinberg: 1992). But while the relativity theory offers an abstract and
quantitative framework for dealing with matter, energy, forces and space, the
operator hierarchy focuses on complementary aspects by offering an abstract
and qualitative framework for organising matter. That the operator hierarchy
deals with qualitative aspects should not be considered a flaw of the theory, but
its strength, because it addresses a blind spot in the scientific literature.
The general analysis of structural hierarchy is not a fashionable topic in science.
Firstly, people may not think about a unifying ranking because they consider
particles, such as hadrons, atoms and molecules, as incomparable with
organisms. Secondly, people may have difficulty identifying a general ranking
rule. When looking at the mechanisms, they look differently at every level. Only
the use of first-next possible closure offers a principle that can be used across
levels. Thirdly, people may consider it wrong to focus primarily on the operators
because the universe is full of interaction systems, such as galaxies, stars,
planets and at least one ecosystem. However, the operator hierarchy cannot be
created or even recognised as long as interaction systems are considered to be
part of its ranking. Fourthly, Teilhard de Chardin’s early work dealing with a
general ranking was confounded with religious statements. This may have led
certain people to reject his early steps towards a general theory altogether.
Finally, the focus of science on quantification and equations has drawn the
attention away from structural analysis, which uses completely different
concepts of quantity. These and other aspects may have contributed to the
absence of the operator hierarchy in any form from the scientific debate.
As was discussed in this study, the operator hierarchy contributes in various
ways to such fundamental topics as a cosmic timeline, an ecological hierarchy, a
198
Chapter 8
periodic table for periodic tables, an organic evolution and an analysis of the
scope of unifying concepts. The operator hierarchy adds to these topics a
unique focus on the structural complexity of systems. This focus enables the
logical integration of distant scientific domains. These achievements support the
conclusion that the operator theory deserves a prominent role in science.
Chapter
General discussion and conclusions
”Der Philosoph ist nicht Bürger einer Denkgemeinde.
Das ist, was ihn zum Philosophen macht.”
(Ludwig Wittgenstein)
9
200
Chapter 9
Introduction
The goal of this thesis was to contribute to unification in science by searching for
rules that guide the development of structural complexity in the universe and that
improve the analysis of hierarchical organisation in biology, ecology and system
science. The major result of pursuing the above goal has been the discovery of
the operator hierarchy. Various individual aspects of the operator hierarchy have
been discussed in the above chapters. To place the conceptual framework of the
operator hierarchy and its applications in a broad context, I will attempt to
answer the following five questions: (1) what kinds of novel theoretical
developments have accompanied the elaboration of the operator hierarchy, (2)
what impacts may the operator theory have on science, (3) what aspects of the
operator theory require discussion because of potential weaknesses, (4) how
can the operator theory be evaluated, and (5) what routes are open for future
development.
What is new?
Three principles are absolutely fundamental to the development of the operator
theory. The first is the reduction of the complexity of the hierarchy problem to its
minimum. The second is the identification of three major dimensions for
analysing hierarchy. The third is the definition of first-next possible closure and
its use as a tool to identify and to rank subsequent operators. To discover
whether these principles can be considered new developments, they will be
individually discussed in relation to existing theory.
Although biologists generally focus on details of a system’s organisation,
physicists quite generally solve problems by reducing the observed complexity
to a minimum. The following citation from Bak (1996) presents an instructive
caricature of the difference in attitude between biologists and physicists:
Thus, how would we physicists make a suitable model of, say,
biological evolution? The biologist might argue that since there is
sexual reproduction in nature, a theory of evolution must necessarily
include sex. The physicist would argue that there was biology before
there was sex, so we don’t have to deal with that. The biologist might
point out that there are organisms with many cells, so we must explain
how multicellular organisms developed. The physicist argues that
there are also single-cell organisms, so we can throw multicellularity
out! The biologist argues that most life is based on DNA, so that
should be understood. The physicist emphasizes that there is simpler
General discussion and conclusions
201
life based on RNA, so we don’t have to deal with DNA. He might even
argue that there must have been a simpler reproductive chemistry
before RNA, so that we don’t have to deal with that either, and so on.
The trick is to stop the process before we throw out the baby with the
bathwater.
That the concept of evolution can be reduced to a minimum shows the
magnificence of the one-line programme developed by Harvey in 1996 (Harvey,
2001, 2009). The programme effectively mimics evolution, including rank
selection, demes, recombination (via horizontal transmission), mutation and
elitism:
for (t=0;t<T;t++) for (W=(e(A=P*r())>e(B=(A+1+D*r())%P)?A:B),
L=(W==A?B:A),i=0;i<L;i++) if ((r=r())<R+M) g[L][i]=(r<R?g[W][i]:g[L][i]^1);
Where: t = tournaments, T = max number of tournaments, W = Winner, L =
Loser, e(i) is a problem-specific evaluation function returning fitness of ith
member of population, g[j][i] is an array giving the ith gene of the jth member of
population, A, B are members chosen for a tournament, to be labelled W(inner),
L(oser) according to which is fitter, r() is a (pseudo-)random number
0.0<=r()<1.0, P is population size, D is deme-size, R is Recombination rate, M is
Mutation rate.
Although the importance of the caricature presented by Bak and the minimalist
model for a biological evolution process designed by Harvey are beyond
discussion, the venom of simplification is in the tail because –as was stated
above- the process should be stopped just in time to prevent the essence of the
problem from being flushed down the drain. It is precisely in relation to this
aspect that the use of first-next possible closure in the operator hierarchy does
represent a major theoretical development; first-next possible closure indicates
the line separating the situation in which the baby is saved from that in which it is
being flushed down the drain. First-next possible closure indicates exactly the
theoretical minimum complexity that is required to form a next level operator. It
does this for all levels in the operator hierarchy, hereby guiding several
situations where the trick is to stop the process of simplification just in time.
The difficulty of stopping the simplification process at the right moment is
illustrated by the common belief that if we understood the origin of life, i.e., the
origin of the first cell, we would solve the long standing problem of the definition
of life. As Jagers op Akkerhuis (2009, in press) has shown, focusing on the first
cells implies that the reduction process has been carried too far. The reason is
that the most detailed insights into a cell’s properties are vastly insufficient to
define even the minimum aspects of all the other complexity levels of life, such
202
Chapter 9
as eukaryotic life, multicellular life, memic life and, in the future, technical memic
life. Using cells as a basis for a definition of life, especially technical life, is
problematic because technical life does not even consist of cells. In the past the
simplification process went wrong because it was continued to the simplest
operator (the cell) and not to the simplest organisation that is typical for every
type of living operator. As a result, all the properties that were typical for the
organisation of complex organisms were lost during the simplification process,
and this loss made it impossible to define life in a general way.
The second major principle advocated by the operator hierarchy was that
organisation should be analysed on the basis of three dimensions for hierarchy.
In principle, this realisation is not new. Around 1940 the idea had been
discussed that when we analyse a system that controls several elements this
system may itself be controlled by another system that resides ‘above’ it
(Feibleman 1954). There are many examples of this three-level approach, for
example, Koestler’s Holon (Koestler, 1967) and Jaros and Cloete’s Biomatrix
(Jaros and Cloete, 1987). It is quite interesting that in many publications such
three-level approaches to systems were placed one after another, such that a
given system with its elements became itself an element in a larger system, etc.
The resulting hierarchy had the effect of guiding the reader’s thoughts towards a
chain of interlocked systems. As a consequence, the initially independent three
levels have now become hostage of a one-dimensional hierarchy. The operator
hierarchy differs on this point because it recognises that hierarchies in these
three levels (interior of a system, the system, and the world) point in different
directions and that for this reason they are independent dimensions for
hierarchy. According to this new perspective, one can analyse hierarchy in the
internal organisation of a system, one can analyse the organisational differences
between systems, or one can analyse how interactions between systems create
hierarchy. Thus, even though the basal three levels of the classical system
theory are retained, the operator theory recognises them as independent
directions for hierarchy and not as levels.
Finally, the use of first-next possible closure in the operator hierarchy seems to
be a truly original theoretical development. Except for Teilhard de Chardin’s
contribution (Teilhard de Chardin, 1969), no publications have been found
dealing with first-next possible closure or related topics referring to “first-next”.
What are the impacts?
The operator theory was developed to solve problems with the integration of
toxicant effects at various levels of ecological organisation. Initially, the
integration was based on the well-known ‘ecological hierarchy’, for example,
from atoms to molecule, organelle, cell, tissue, organ, organism, population,
General discussion and conclusions
203
community and ecosystem, planet, solar system, galaxy and universe. But this
hierarchy proved to be conceptually inconsistent and insufficient. Because no
alternative was found in the literature, the only solution was to try to develop a
new ecological hierarchy. This resulted in the discovery of the operator
hierarchy. Later, it appeared that the logic of the operator hierarchy could be
applied to solve many other problems that had their origin in the inconsistent use
of hierarchy. Most notably, the operator hierarchy was applied to rearranging the
classical ecology hierarchy. By using the operator theory, the original linear
ranking of elements in this hierarchy could be transformed into three separate
rankings: one for the hierarchies in interaction systems, one for the hierarchy of
the operators and one for hierarchies in the internal organisation of the
operators.
Furthermore, the operator hierarchy has a fundamental application in arranging
a happy marriage between the particle concept used in physics and chemistry
and the organism concept used in biology and ecology. Based on the operator
hierarchy, the two concepts were united in a single ranking in which the word
‘operator’ was used as a common concept. Integrating all operators in one
ranking also changes the way in which a particle is defined. Although the
classical concept of a particle concentrates on minute physical entities (for
example, see Falkenburg’s review, 1993) one must now consider that the
particle concept can be broadened to the operator concept and that demands for
first-next possible closure should be part of a definition of physical and chemical
particles.
Because it ranks abiotic and biotic operators, the operator hierarchy can also be
applied as a periodic table of periodic tables because for every operator -- and
for some interfaces or pre-operator hypercyclic sets -- a corresponding periodic
table exists for the classification of all the known operators at that complexity
level. In this way, the operator hierarchy connects the standard model for
fundamental particles, the eightfold way for hadrons, the periodic table of the
elements and the tree of life.
Another application of the operator hierarchy was the further development of
complexity layers in the tree of life. In its most basal representation, the tree of
life presents the historical divergence of species. Quite frequently, the time axis
is supplemented with information about whether the organisms are prokaryotic,
eukaryotic or multicellular. The operator hierarchy now suggests adding two
other classes: multicellular prokaryotes (e.g., blue-green algae) and memons.
The major advantage of introducing memons is that brains are no longer
regarded as just another organ that can be found in multicellular organisms but
as a different level of organisation altogether. The recognition of a level of
memic organisation also has the advantage of fully merging neurobiology and
technical intelligence.
204
Chapter 9
The operator hierarchy also offers an interesting solution to the definition of life
problem. Earlier attempts at a definition frequently focused on the origin of life,
for example, on the properties that were typical for the minimum organisation of
the most primitive prokaryote cell. The operator hierarchy indicates however,
that minimum organisation is not typical for prokaryotes but exists at many
levels, including those of the eukaryotes, the multicellulars, and the memons.
Accordingly, a definition of life cannot be complete if it does not consider the
different minimum organisations at all these levels. Using the ranking of the
different operator types offered by the operator hierarchy, life is defined as
matter with the construction of operators under the limitation that the operators
should at least show the cellular operator’s complexity. By focusing on operators
and their minimal construction properties, the operator-based definition also
solves problems with facultative aspects that plague many other definitions. For
example, frozen cells are still life, according to the operator-based definition,
even though they do not show physiology, behaviour, etc. Similarly, reproduction
and growth are no longer requirements for life, although the production of
offspring is, of course, required for evolution! Another implication of the operatorbased definition is that technical memons must be regarded as life because it is
irrelevant, for example, whether the structure of hardwired memons is based on
organic cells or on technical devices.
A field of theory where the operator hierarchy has also been applied is that of
astrobiology and extraterrestrial life. If it is assumed that the operator hierarchy
correctly describes all the first-next possible closures that are possible in our
universe, then the same operators will be formed anywhere in the universe for
as long as ambient conditions allow this to happen. This assumption is
supported by observations showing that atoms and molecules also exist in
distant galaxies. Similarly, life, as defined by the operator hierarchy, may exist in
the universe on other planets than Earth.
Yet another application of the operator hierarchy is that it offers a structured way
to reason about the construction of future operators and hence about the future
of evolution and, in particular, artificial intelligence (AI). Even when the outcome
may be the same, using the operator hierarchy gives predictions a better basis
than what is offered by observing trends in the robotic industry. Of course, the
extrapolation of the operator hierarchy holds only as long as the sequence of the
operators and their secondary structure are interpreted correctly. Under these
assumptions, the extrapolation of the operator hierarchy can be used to predict
the closures of a range of future operators. Although the operator hierarchy may
well offer the only method science has ever had to predict future particles, the
result remains, of course, an extrapolation and is burdened with various
uncertainties. The most important uncertainty is related to how a closure
dimension should be interpreted at a higher level. For example, the distance
between multiatomarity and multicellularity is already enormous, but the distance
General discussion and conclusions
205
between multicellularity and multimemicity is even larger. Moreover, where it is
already difficult to predict details of multicellularity as long as only bacteria are
known, it will be much more difficult to predict details of multimemic life if the
only known memons are animals. Of course, the smaller the predicted steps
forward, the more likely it will be that a prediction is correct. The smallest step
that can be predicted using the operator hierarchy is that from the hardwired
memon (animals with ‘brains’, including human beings) to the softwired memon
(a technical life form with modelled brains). The operator hierarchy predicts that
a memic architecture is required if technical devices should show the emergent
properties typical for memons, such as learning, association, creativity,
intelligence, thinking and consciousness. These properties are all graded
properties that increase in close correlation with the complexity of the neural
network (e.g., size, connectivity, managing of weights of neural connections,
hierarchy of feedback loops, etc.). In addition, a neural network has the
advantage of efficiency because it does not have to store all past perceptions in
files whose content has to endlessly increase over time with every experience
(Baxter and Browne, 2009). Instead, a neural network stores new information by
fine-tuning the structuring of one and the same network.
The operator hierarchy could furthermore be applied in the field of artificial
intelligence to redefine the meme concept. The classical meme concept was
introduced by Dawkins (1989). Dawkins defines the meme as a concept that
“conveys the idea of a unit of cultural transmission or a unit of imitation”. The
meme has a deliberate association with the gene concept that represents the
unit of heritable transmission. However, there is no proper parallel between the
two concepts. Genes are material structures that code for organism properties.
Memes are concepts of cultural transmission without a clear structural basis.
The operator hierarchy suggests that the meme concept can be given a
structural basis, if memes are defined as the coding that in technical memons
(the softwired memons) contains the information about the number of neurons,
their connections and the strength of these connections (Jagers op Akkerhuis,
2001).
The operator theory also offers a context for discussing the arrow of complexity
hypothesis. On the arrow of complexity hypothesis, Bedau (2009) states the
following:
We can distinguish three things: (i) a trend, which is simply a
directional change in some variable; (ii) a robust regularity, which is
a generic or non-accidental trend and which can be thought of as a
statistical “law” of nature; and (iii) a mechanism which explains how
a trend (whether accidental or robust) is produced. Now we can
formulate the arrow of complexity hypothesis—the hypothesis that
evolution inherently creates increasingly complex adaptive
206
Chapter 9
organisations as life forms. Note that the hypothesis is about the
increasing complexity of the most complex forms of life, not of all
life forms. The hypothesis concerns not mere complexity but
adaptive complexity. The arrow hypothesis should be understood
as a robust regularity that has exceptions, and it holds for evolving
systems only given certain constraints.
Assuming that both a trend and a robust regularity require a mechanism that
explains their occurrence, confirmation of the arrow of complexity hypothesis
requires an explanatory mechanism that, in a robust and regular way, increases
the complexity of the most complex life forms. Demonstrating such a mechanism
implies proof of the evolutionary advantages of complexity and a yardstick to
measure it. The operator hierarchy solves the last problem by offering a strict
yardstick as the number of closures a system shows and the elaborateness of a
closure when present. Accordingly, plants are more complex than protozoa and
single-celled blue-green algae are more complex than Pseudomonas bacteria.
An indirect proof for the evolutionary advantage of complexity can be derived
from the fact that no system can organise itself in contradiction to the laws of
thermodynamics. Accordingly, the fact that high complexity memic operators
exist on Earth implies that complexity must have advantages that outweigh the
costs. Other approaches to prove the advantage of high complexity may involve,
for example, the economy and the competitive power of three-dimensional
multicellular shapes compared to those of single-celled organisms. However, if it
is assumed that the operator hierarchy correctly indicates the only closure
situations that are possible, this assumption implies a fixed direction for the
arrow of complexity because any evolution process either comes to a halt or
proceeds along the pathway of the operator hierarchy.
What are potential weaknesses?
The operator hierarchy was constructed with great care. Every step in the
hierarchy was checked to confirm if no other layer could be added or if a layer
could be made superfluous. In addition, the internal logic of the operator
hierarchy based on closure dimensions and complexity layers acts as a control
mechanism because it leaves little freedom for manipulation. Its applications
show that the idea is compatible with a whole range of other scientific ideas that
could be supplemented or improved by relating them to the operator hierarchy.
Despite these positive aspects, it is important to remain critical and to
continuously search for ways to falsify or improve the theory. The aim of this
chapter is therefore to critically discuss various aspects of the operator hierarchy
that may lead to questions or that require further research.
General discussion and conclusions
207
First of all, it must be said that the operator hierarchy has a strict and logical
structure but that there is no clear explanation for the fact that every primary
operator adds a new closure dimension. Although this may have something to
do with the increasing conformational richness of every next hypercyclic
configuration lying at the basis of every primary operator, it is still unclear how
this mechanism works. It is easier to explain the observation that every primary
operator shows transitions that recreate every existing closure dimension. The
reason for this is that the transition cannot be made in another way (e.g., from
quark to hadron, from atom to molecule) or that all possible pathways are in
principle explored (e.g., from prokaryote unicellular to eukaryotic unicellular to
prokaryotic and eukaryotic multicellular).
Secondly, the operator hierarchy is based on hypothetical closures that
represent the least complex configurations. Consequently, the closures that
figure in the operator hierarchy will generally have no parallels in real life. Even
the simplest bacterial cell will most likely require more than two catalysts to
perform autocatalysis while maintaining the cell membrane. This artefact results
from the intention to minimise complexity in order to gain more insight into what
are the essential aspects of every transition in the hierarchy.
Thirdly, it can be suggested that “”there are more things in heaven and earth,
Horatio, than are dreamt of in your philosophy” (William Shakespeare, Hamlet:
Act 1. Scene V). In fact, this is a most interesting criticism. From science fiction
stories, astrology programmes on television, biblical assumptions about
resurrection and life after death, fairytales, ghost stories and the general
scientific message that many things are still waiting to be discovered, most
people are convinced that there is still much we do not know. The operator
hierarchy offers a scientific way to discuss this point of view. The hierarchy’s
logic implies that the space for potential novelty is rather limited. The reason is
that the operator hierarchy ranks all the operator types that have occurred since
the beginning of the universe from the fundamental particles to the memons.
Furthermore, for all operator types, science has created and continues to create
increasingly detailed insight into what forms the operators may take, how they
interact with their environment and what they may transform into (when they
can). Moreover, assuming that the operator hierarchy ranks all possibilities for
first-next possible closure, no other types of operators can exist elsewhere in the
universe. Accordingly, the space that is left for truly unexplainable things has
become small indeed. That the space for unexplainable things is limited has also
been convincingly advocated by ‘t Hoofd (2000). Of course, a reasoning like that
given above is not complete without indicating that science can say little about
things that it has not measured or cannot measure. How much space is left for
new discoveries depends in the end on whether a natural limit exists to the
number of phenomena that can exist in our universe at the different levels of
organisation and as the result of all possible interactions between elements.
208
Chapter 9
Fourthly, it may be argued that the danger of a scientific Procrustusian bed is
lurking in the operator hierarchy. As an old Greek story goes, Procrustus was a
blacksmith who invited people to sleep in his house. To a tall person he offered
a small bed, and while the person slept, Procrustus hammered off his guest’s
extremities that extended beyond the bed. Similarly, a short person was offered
a large bed, and Procrustus stretched his body by hammering on it until it fit the
bed. In the context of debunking the operator theory, it is a legitimate question
whether the notion of a rough blueprint of the operator hierarchy has not been a
reason for formulating definitions that are fine-tuned to confirm the theory while
a more general application of the principles would have led to a different
outcome. This is an unlikely scenario because the definitions that describe the
operator hierarchy were constructed so that the mechanisms involved are as
general as possible to avoid manipulating definitions towards specific situations.
Furthermore, any set of definitions that has been adapted to a Procrustusian bed
will fail to fit with other approaches that are free of such limitations. In this sense
it can be regarded as positive circumstantial evidence that the operator
hierarchy has led to useful applications in various scientific fields (e.g.,
ecological hierarchy, definition of life).
Fifthly, it must be concluded that the operator hierarchy offers no information
about the average time that is required to reach any next level in the operator
hierarchy. For this reason, the operator hierarchy lacks direct associations with
time-related topics, such as evolutionary acceleration; it has a strict focus on
structural transitions. The absence of time is not a unique property of the
operator hierarchy. Time is not quantified in either thermodynamics or Darwinian
evolution. But even though the operator hierarchy gives no information on this
point, there are several reasons why an evolutionary process may speed up
over time.
a. Development acceleration
If elements remain in the system once they have been formed, this will
increase the possibilities for novel interactions. If we now postulate an
evolutionary rate as the chance per time that new elements are formed, this
rate will actually represent acceleration because the evolutionary rate creates
new elements on the basis of the cumulated number of elements formed. The
combination of both processes implies that the evolutionary rate (dimension
time-1) transforms into an evolutionary acceleration (dimension time-2).
b. Modularity and hierarchy
Assuming a system of k building block elements that shows a hierarchical
structure with many layers, and assuming that on average s components
have to combine to form the elements of any next level in the hierarchy and
take the same time to do so, it can be deduced that the time for the evolution
of the whole system is proportional to the logarithm to base s of k. Given any
system that is structured in this way, the time required to develop from a
General discussion and conclusions
209
system with 1 element to a system with 1000 elements would be the same as
for a system of 1000 elements to reach the size of a million elements (see
also Simon, 1962).
c. The evolution of evolvability
Organisms are selected for a combination of complexity and efficiency that
combines resource dominance strategies that, in a given environment, yield
the highest increase in resource dominance. The organisms success in these
combined efforts, e.g., their evolutionary success, can be measured either in
terms of their present resource dominance or in terms of the increase and the
acceleration hereof. A focus on these different rates is of marked importance;
those that are the best now may soon lose to those that slowly improve and
both strategies will be beaten in the long term by units that can increase their
improvement rate. Early papers on the importance of the evolution of
evolvability have been published by Wagner (1996).
d. The use of larger energy gradients in combination with increased efficiency
This aspect selectively applies to living operators. To maintain their
organisation, organisms need energy. If there is a large quantity of energy
available, the organism can afford a larger organisation. For this reason,
increasing complexity requires increasing efficiency and/or an increase in
access to energy gradients. In this sense, the changes from litho-chemical
metabolism to photosynthesis, to the consumption and digestion of organic
material from other organisms, to fossil fuel, nuclear energy and solar energy
can be regarded as innovations that have sustained and/or accelerated the
evolution process.
e. The evolution of concepts
Once the level of memons has been reached, new discoveries can be
learned and/or documented. In this way, knowledge can be disseminated
(also without reproduction) and be used to create new tools to further
increase the rate of new discoveries and the development of knowledge. This
acceleration mechanism must be considered the cause of the knowledge
explosion crises.
Integrating the above and possibly other processes, Kurzweil (1999) has
proposed the “law of time and chaos”, as follows: ‘In a process, the interval
between salient events (that is, events that change the nature of the process, or
significantly affect the future of the process) expands or contracts along with the
amount of chaos’. Clearly, ‘time’ in this viewpoint is relative to the frequency of
salient events. From this general viewpoint, two laws that are more specific can
be derived, depending on whether one considers increasingly chaotic systems
or increasingly ordered systems. For increasingly chaotic systems, like the
overall universe, the ‘negative’ version of the above law – ‘The law of increasing
chaos’ -- reads as follows: ‘As chaos exponentially increases, time exponentially
slows down; that is, the time interval between salient events grows longer as
210
Chapter 9
time passes’. This fits well to the processes that can be observed in the overall
universe, where most salient events, such as the formation of matter, occurred
in the first seconds of the universe’s existence. The ‘positive’ version of the
above law – ‘The law of accelerating returns’-- on the other hand, can be
formulated as follows: ‘As order exponentially increases, time exponentially
speeds up; that is, the time interval between salient events grows shorter as
time passes’. The latter fits well with the acceleration of evolution observed in
local spots of the universe where initial self-organisation processes (e.g.,
chemistry, autocatalysis) lead to increasingly rapid self-organisation (e.g., cells,
multicellulars, memons, etc.). Together these two laws describe that in the
universe, the overall processes increase chaos and slow time, whilst in some
local systems order has increased and continues to produce more order in an
increasingly rapid way.
The law of time and chaos uses the word ‘salient’ events. I consider this
formulation in perfect harmony with the operator theory because the operator
hierarchy describes the occurrence of specific emergent properties that lead to
the construction of new operators. The transitions leading from one operator to
the next typically signify the most important ‘salient’ events in the universe.
Of course, the above examples of acceleration are simplified generalisations
that could suffer from various disturbances. For example, certain newly formed
elements may negatively affect the formation of new elements by disrupting the
system. Moreover, adding higher level elements may block certain types of
interactions with lower levels, hereby reducing the number of interactive
combinations. It may also be argued that at any moment in evolution, the
possibility of forming new elements from the prevailing elements has simply
been zero (e.g., there is no life on the moon), which makes the evolutionary rate
drop to zero. Yet, the long array of system evolution from quarks to neural
networks shows that, at least in our part of the universe, a fatal blocking of the
process has not yet occurred.
Sixthly, the question arises whether the operator theory represents a falsifiable
construction, a construction in which experiments can be designed to prove or
disprove the theory. The operator theory is, however, not a quantitative theory. It
deals with the recognition and ranking of system configurations resulting from
self-organisation processes. Is it possible to falsify such an approach? In this
respect the following points may be considered to test the theory: (1) Is the
ranking strict? (2) Is the second-order ranking by closure dimensions consistent?
(3) Do any of the proposed transitions contradict thermodynamic laws? (4) Can
the ranking be used to predict gaps? Although nothing indicates that the
operator theory would violate the first three of these aspects, it is scientifically
desirable to search for ways that can prove or disprove these aspects in a
definite way. A difficulty with respect to the fourth point is that the theory cannot,
General discussion and conclusions
211
in principle, show gaps because it is designed as a strict ranking. If there were
gaps, this would immediately require another ranking. The prediction of gaps is
therefore limited to future operators that are not yet included in the ranking. Only
the future will tell whether the predictions based on the operator hierarchy are
correct.
From the above points it can be concluded that certain aspects of the operator
theory need to be further developed. The identification of operators showing the
same closure dimension would especially profit from a less intuitive conceptual
methodology. On the other hand, no aspects have been found where the
operator hierarchy can easily be proven wrong or insufficient.
How can the operator theory be evaluated?
When it comes to judging the value of the operator theory, certain evaluation
criteria are needed. In a review of theories for worldview comparison, Vidal
(2009) offers a set of three objective criteria to evaluate and compare
worldviews: (1) the theory should exhibit internal and systemic consistency, (2)
the theory should fit with all natural sciences, and (3) the theory should address
and adequately resolve worldview questions. Vidal mentions the following
worldview questions in his paper: What is? Where does it all come from? Where
are we going? What is good and what is evil? How should we act? What is true
and what is false?
Vidal’s criteria offer guidance to objectively discuss the operator theory’s value
(Vidal 2009). First of all, the theory should exhibit internal end systemic
consistency. In this respect, it may be judged as positive that all steps in the
operator hierarchy are based on first-next possible closures and that a set of
coherent definitions was used for its creation. A further indication for a strong
internal consistency comes from the fact that no single operator can be taken
from or added to the operator hierarchy without destroying the entire framework.
Secondly, the operator hierarchy should fit all natural sciences. It should be
stressed that the use of sound scientific criteria has always been the number
one criterion for constructing the operator theory. All transitions and all system
types have been checked – as much as possible -- for being based on testable
scientific concepts. Thirdly, the theory should address and adequately resolve
worldview questions. Although I am not sure whether other people might
suggest additional worldview questions, the operator theory addresses at least
three of Vidal’s questions. The question of “What is?” is addressed by the
operator hierarchy, because the hierarchy hypothesizes that all structures in the
universe are based on one of the system types in the operator hierarchy
(including the pre-operator hypercyclic sets and the interfaces). While unifying
theories such as the grand unification theory in physics address particles and
212
Chapter 9
forces at one level of organisation, the operator hierarchy addresses particles at
all levels of organisation. As it addresses more levels, the operator hierarchy can
be used to suggest an answer to the questions of “What is life?” and “What is
death?”. The question “Where does it all come from?” is partially addressed by
the long sequence of operators illustrating the major organising principles that
allowed the successive system types to come into being. Finally, the operator
hierarchy offers a unique possibility to answer the question of “Where are we
going?” because it predicts the major structural aspects of future operators, from
which information can be gained about their interactive properties and thus, how
they might become players in the world that we still consider ours. In fact, the
operator theory can also be considered to address the question of “What is true
and what is false?” because it has helped improve various definitions, including
the definition of a particle, life, the organism, death and ecological hierarchy.
Future development
I am convinced that the most promising routes for future development lie in
finding explanations for the second-order structure of the operator hierarchy.
Insights into this structure will test the validity of the operator hierarchy. If the
operator hierarchy’s basic structure can be consolidated and our understanding
deepened, I see three exciting routes for future development. The first is to
increase insight into the lowest levels of organisation. The second is to apply the
operator hierarchy’s structure to solving long standing problems related to
hierarchical organisation. The third is to increase both the detail and the reach of
predictions of future operators.
In conclusion
Originally, the search for order that is described in this thesis started with an
integration problem in ecotoxicology. But the direction of the project changed
when the first ideas about the operator hierarchy emerged; it suddenly became
possible to pursue the much broader goal of contributing to the unification in
science by searching for rules that guide the development of structural
complexity in the universe. As was demonstrated in the different chapters of this
thesis, the operator hierarchy holds a central position in the organisation of
matter. The strict ranking of the operators based on first-next possible closure
offers a basis for several important applications, including the definition of life,
the restructuring of the classical ecological hierarchy, the prediction of future
operators and the offering of a scientific integration tool in the form of a periodic
table for periodic tables. In this way, the operator hierarchy improved the
analysis of organisation in biology, ecology and system science. Given its
central position in natural organisation and its broad range of fundamental
General discussion and conclusions
213
applications, the operator hierarchy can be regarded as a unique step forward
towards Von Bertalanffy’s dream of a ‘general theory of hierarchic order’ (Von
Bertalanffy, 1968) .
214
Chapter 9
References
216
References
Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K. & Watson, J. D., 1989. Molecular
biology of the cell. 2e ed. Garland Publishing Inc. New York.
Allen, J.F., 1993. Control of gene expression by redox potential and the requirement for
chloroplast and mitochondria genomes. J. of Theor. Biol. 165: 609-631.
Altmann, R., 1890. Die Elementarorganismen und ihre Beziehungen zu den Zellen (Veit,
Leipzig, 1890).
Alvarez de Lorenzana, J.M. & Ward, L.M., 1987. On evolutionary systems. Behav. Sci.
32:19-33.
Alvarez de Lorenzana, J.M., 1993. The constructive universe and the evolutionary
systems framework. (In: S.N. Salthe,1993. Development and evolution.
Complexity and change in biology. MIT Press, Cambridge. Appendix pp 291-308)
Arditi, R., Michalski, J. & Hirzel, A.H., 2005. Rheagogies: modelling non-trophic effects in
food webs. Ecological Complexity 2: 249-258.
Aristotle, Metaphysics. Book H, 1045:8-10
Bak, P., 1996. How nature works: The Science of Self-Organized Criticality. Copernicus,
Springer-Verlag, New York.
Bardele, C.F., 1997. On the symbiotic origin of protists, their diversity, and their pivotal
role in teaching systematic biology. Ital. J. Zool. 64: 107-113.
Baxter, P., Browne, W., 2009. Memory-Based Cognitive Framework: a Low-Level
Association Approach to Cognitive Architectures. ECAL conference 2009,
Budapest.
Becquerel, P., 1950. La suspension de la vie des spores des bactéries et des
moisissures desséchées dans le vide, vers le zéro absolu. Ses conséquences
pour la dissémination et la conservation de la vie dans l’univers. Les Comptes
rendus de l’Académie des sciences, Paris, 231: 1274 & 1392-1394.
Becquerel, P., 1951. La suspension de la vie des algues, lichens, mousses, aux zéro
absolu et role de la synérèse réversible pour l’existence de la flore polaire et des
hautes altitudes. Les Comptes rendus de l’Académie des sciences, Paris, 232: 2225.
Bedau, M.A, 2009. The evolution of complexity. In: Barberousse, A., Morange, M., and
Pradeu, T. (Eds.): Mapping the Future of Biology. Boston Studies In The
Philosophy Of Science, Volume 266, Springer Netherlands.
Bedau, M.A., 2007. What is life. In: S. Sarkar & A. Plutynski, eds. A companion to the
philosophy of biology, pp. 455-603.
Bejan, A. & Lorente, S., 2006. Constructal theory of generation of configuration in nature
and engineering. Journal of applied physics 100 (041301; doi:10.1063/1.2221896)
Bejan, A., Lorente, S., 2004. The constructal law and the thermodynamics of flow
systems with configuration. International journal of heat and mass transfer 47:
3203-3214.
Bejan, A. & Marsden, J.H., 2009. The constructal unification of biological and geophysical
design. Physics of Life Reviews 6: 85-102.
Beloussov, L.V, 1993. Transformation of morphomechanical constraints into generative
rules of organic evolution. World Futures 38: 33-43.
References
217
Benda, C., 1898. Ueber die Spermatogenese der Vertebraten und höherer Evertebraten,
II. Theil: Die Histiogenese der Spermien. Arch. Aat. Physiol. 73: 393-398.
Bernard, C., 1865. Introduction à l’étude de la médecine expérimentale, Nombreuses
rééditions, p. ex.: 1943, Paris, Librairie Joseph Gibert (col. ‘Chefs-d’oeuvre
philosophiques’, 308: 175.
Berg, O.G. & Kurland, C.G., 2000. Why mitochondrial genes are most often found in
nuclei. Molecular Biology and Evolution 17: 951-961.
Bergson H., 1911. Creative evolution. New York: Holt and Co.
Bertalanffy, L. von, 1950. An outline of general system theory. The British Journal for the
Philosophy of Science. 1: 134.
Bertalanffy, L. von, 1968. General System Theory, foundations, development,
applications. Pinguin Books Ltd. Harmondsworth, Middlesex, England.
Bertalanffy, L. von, 1968. General systems theory. George Braziller, New York.
Blackmore, S., 1999. The meme machine. Oxford University Press.
Blackstone, N.W., 1995. A units-of-evolution perspective on the endosymbiont theory of
the origin of the mitochondrion. Evolution 49: 785-796.
Bonner, J.T., 1998. The Origins of Multicellularity. J.T. Bonner. Integrative Biology 1: 2736.
Bro, P., 1997. Chemical reaction automata. Precursors of artificial organisms. Complexity
2: 38-44.
Broca, P., 1860-1861. Rapport sur la question soumise à la Société de Biologie au sujet
de la reviviscence des animaux desséchés. Mém. Soc. Biol., 3me Série, II, 1860:
1-139.
Bryson, B., 2003. A short history of nearly everything. USA: Broadway Books.
Bullock, S., Noble, J., Watson., R. & Bedau, M. A. (eds.), 2008, Artificial Life XI:
Proceedings of the Eleventh International Conference on the Simulation and
Synthesis of Living Systems. MIT Press, Cambridge, MA.
Bunge, M., 1969. The metaphysics, epistemology and methodology of levels. In
‘Hierarchical structures’ (L.L. Whyte, A.G. Wilson and D. Wilson, Eds.), pp. 17-26.
Am. Elsevier, New York.
Bunge, M., 1992. System boundary. Int. J. Gen. Syst. 20: 215-219.
Calvin, W.H., 1987. The river that flows uphill A Journey from the Big Bang
to the Big Brain. Sierra Club Books.
Campbell, D.T., 1960. Blind variation and selective retention in creative thought as in
other knowledge processes. Psychological Review 67: 380-400.
Campbell, D.T., 1990. Levels of organization, downward causation, and the selectiontheory approach to evolutionary epistemology. In: Greenberg, G. and Tobach, E.
(Eds.), Theories o the evolution of knowing, T.C. Schneirla Conference Series, vol.
4, Hillsdale, NJ: Erlbaum, 1-17.
Capps, G.J., Samuels, D.C. & Chinnery, P.F., 2003. A model of the nuclear control of
mitochondrial DNA replication. Journal of Theoretical Biology 221: 565-583.
Capra, F, 1996. The web of life: a new scientific understanding of living systems. Anchor
Carnot, S., 1824. Réflexions sur la puissance motrice du feu et sur les machines propres
à développer cette puissance. Paris: Bachelier.
218
References
Chadwick, J., 1932. The Existence of a Neutron. Proceedings of the Royal Society,
Series A, 136: 692-708
Chaisson, E.J. 2001. Cosmic Evolution: The Rise of Complexity in Nature Harvard
University Press, Cambridge, 2001.
Chandler, J. L. R. & Van de Vijver, G., 2000. Eds.: Closure: emergent organisations and
their dynamics. Annals of the New York Academy of Sciences 901.
Checkland, P. & Scholes, J., 1990. Soft systems methodology in action. John Wiley &
Sons, Chichester, 329 pp.
Claessens, M.M.A.E., Leermakers, F.A.M., Hoekstra, F.A. & Cohen Stuart, M.A., 2007.
Entropic stabilisation and equilibrium size of lipid vesicles. Langmuir 23: 63156320.
Clark, C.G., Roger, A.J., 1995. Direct evidence for secondary loss of mitochondria in
Entamoeba hitolytica. Proceedings of the National Academy of Sciences of the
USA 92: 6518-6521.
Clausius, R., 1865. The Mechanical Theory of Heat – with its Applications to the Steam
Engine and to Physical Properties of Bodies. London: John van Voorst, 1
Paternoster Row. MDCCCLXVII.
Cleland, C.E. & Chyba, C.F. 2002. Defining ‘life’. Origins of life and evolution of the
Biosphere 32: 387-393.
Cleland, C.E. & Chyba, C.F., 2007. Does ‘life’ have a definition? In: Planets and Life: The
Emerging Science of Astrobiology, Eds. Woodruff, T., Sullivan, I.I.I., Baross, J.A.
Cambridge University Press.
Close, R, 1983. The cosmic onion. Quarks and the nature of the universe. Heinemann
Educational Books Ltd. USA.
Conrad, M., 1982. Bootstrapping model of the origin of life. Biosystems 15: 209-219.
Cohen, J. & Stewart, I., 1994. The Collapse of Chaos: Discovering Simplicity. Viking,
Penguin, London.
Curie, M. P., Mme. Curie P. & Bémont, M. G. 1966. Presented by M. Becquerel: On a
new, strongly radioactive substance, contained in pitchblende, note by Comptes
Rendus 127: 1215-7 (1898) translated and reprinted, In: Henry A. Boorse and
Lloyd Motz, eds., The World of the Atom, vol. 1 (New York: Basic Books, 1966)
Dalton, J., 1808. A new system of chemical philosophy. 3 vols. Manchester, 1808, 1810,
1827.
Darwin, C., 1859. On the origin of species by means of natural selection. Reprinted.
London: Penguin.
Dawkins, R., 1976. The selfish gene. Oxford University Press, Oxford.
Dawkins, R., 1996. Climbing mount improbable. Viking.
Dittrich, P. & Speroni di Fenizio, P., 2007. Chemical organisation theory. Bulletin of
mathematical biology 69: 1199-1231.
Doolittle, W.F., 1998. A paradigm gets shifty. Nature 392: 15-16
Eck, R., & Dayhoff, M.O., 1966. Evolution of the structure of ferredoxin based on living
relics of primitive amino acid sequences. Science 152: 363-366.
Edwards, M. R., 1998. From a soup or a seed? Pyritic metabolic complexes in the origin
of life. Trends in Ecol. Evol. 13: 178-181.
References
219
Eigen, M., 1971. Molekulare Selbstorganisation und Evolution (Self organisation of matter
and the evolution of biological macro molecules.) Naturwissenschaften 58: 465523.
Eigen, M., & Schuster, E., 1977. The hypercycle: a principle of natural self-organisation,
Part A: The emergence of the hypercycle. Naturwissenschaften 64: 541.
Eigen, M. & Schuster, P., 1978a. The hypercycle: a principle of natural self organisation,
Part B: The abstract hypercycle. Naturwissenschaften 65: 7-41.
Eigen, M. & Schuster, P., 1978b. The hypercycle: a principle of natural self-organisation,
Part C: The realistic hypercycle. Naturwissenschaften 65: 341-369.
Eigen, M. & Schuster, P.,1979. The hvpercycle: a principle of selforganizalion. Springer,
New York.
Eigen, M. & Winkler, R., 1983. Laws of the Game, How the principles of nature govern
change. New York: Harper and Row.
Eigen, M., 1985. Macromolecular evolution: Dynamical ordering in sequence space. In
Pines D. (Ed.): Emerging synthesis in science: Proceedings of the founding work
shop of the Santa Fe Institute, Santa Fe, N.M.
Eldredge, N., 1985. Unfinished synthesis. Biological hierarchies and modern evolutionary
thought. Oxford University Press, New York.
Emmeche, C., 1997. Autopoietic systems, replicators, and the search for a meaningful
biological definition of life. Ultimate Reality and Meaning 20: 244-264.
Etxeberria, A., 2004. Autopoiesis and natural drift: genetic information, reproduction, and
evolution revisited. Artificial Life 10: 347-360.
Falkenburg, B., 1993. Was ist ein Teilchen? Bedeutungen eines fundamentalen
physicalischen Konzepts. Phys. Bl. 49: 403-408.
Fanelli, D. & McKane, A.J., 2008. Thermodynamics of vesicle growth and instability.
Physical Review E 78: 1-9.
Fear, K.K. & Price, T., 1998. The adaptive surface in ecology. Oikos 82: 440-448.
Feibleman, J.K., 1954. Theory of integrative levels. British Journal Phylosophy Sci. 5: 5966.
Feynman, R.P., 1985. QED: The strange theory of light and matter. Princeton: Princeton
University Press.
Frank, S.A., 1997. The price equation, Fisher's fundamental theorem, kin selection, and
causal analysis. Evolution 51: 1712-1729.
Gánti, T., 1971. The principle of life (in Hungarian). Gondolat, Budapest.
Gánti, T., 2003a. Chemoton Theory Vol. 1. Theoretical Foundation of Fluid Machineries.
Kluwer Academic/Plenum Publishers, 2003
Gánti, T., 2003b. Chemoton Theory Vol. 2. Theory of Living Systems. Kluwer
Academic/Plenum Publishers, 2003 .
Garnier, C., 1897. Les filaments basaux des cellules glandulaires. Note préliminaire.
Bibliogr. Anat. 5: 278-289.
Gavrilets, S., 1997. Evolution and speciation on holey adaptive landscapes. - Trends
Ecol. Evol. 12: 307-312.
Geiger, G., 1990. Evolutionary instability: logical and material aspects of a unified theory
of biosocial evolution. Springer-Verlag, Berlin.
220
References
Geiger, H. & Marsden, E., 1913. LXI. The Laws of Deflexion of a Particles through Large
Angles. Philosophical Magazine Series 6, Volume 25, Number 148
Geiger, H. & Marsden, E., 1909. "On a Diffuse Reflection of the α-Particles". Proceedings
of the Royal Society, Series A 82: 495–500.
Gell-Man, M. & Neeman, Y., 1964. The eightfold way. W A. Benjamin, Inc., NY,
Amsterdam.
Gilbert, T.L., 1983. Unifying physical concepts of reality. Presentation for the Science
Symposium Session at the second biennial meeting of the Institute for ultimate
Reality and meaning, Toronto, august 1983.
Gentner, D. & Jeziorski, M., 1993. The shift from metaphor to analogy in Western
science. In: A. Ortony (Ed). Metaphor and thought (2nd ed.) pp. 447-480. New
York, NY, US: Cambridge University Press. xvi, 678 pp.
Gleick, J., 1987. Chaos: Making a New Science. Viking, Penguin.
Gödel, K., 1938. The consistency of the axiom of choice and of the generalized coninuum
hypothesis. Proc. Nat. Acad. Sci. USA , 24: 556–557.
Goguen, J. & Varela, F., 1979. Systems and distinctions, duality and complementarity.
Int. f. Gen. System 5: 31-43
Gould, S.J., 1989. Wonderful Life: The Burgess Shale and the Nature of History, Norton,
New York (1989).
Gray, M. W., 1992. The endosymbiont hypothesis revisited. Int Rev Cytol 141: 233-357.
Greben, J.M., 2010. The Role of Energy Conservation and Vacuum Energy in the
Evolution of the Universe. Foundations of Science 15: 153-176.
Grew, N., 1682. Anatomy of plants. London
Gribbin, J. & Gribbin, M., 2008. The universe, a biography. Penguin books.
Grosberg, R.K. & Strathmann, R.R., 2007. The evolution of multicellularity: A minor major
transition? Annual Review of Ecology, Evolution, and Systematics 38: 621-654.
Gutierrez, A.P., 1996. Applied population ecology. A supply-demand approach. - John
Wiley, New York.
Guttman, B.S., 1976. Is ‘Levels of organisation’ a useful biological concept? Bioscience
26: 112-113.
Haber, W., 1994. System ecological concepts for environmental planning. In: Ecosystem
classification for environmental management (ed. E Kleijn), pp. 49-67. Kluwer Ac.
Publ.
Hamilton, W.D., 1964. The genetical evolution of social behavior. J. Theor. Biol. 7: 1-52.
Happel, B.L.M. & Murré, J.M.J., 1994. Design and evolution of modular neural network
architectures. Neural networks 7: 985-1004.
Happel, B.L.M., 1997. Principles of neural organisation: Modular neuro-dynamics. PhD
thesis, 125 pp.
Hawking, S., 1988. A brief history of time. New York, Bantam.
Harvey, I., 2001. Artificial Evolution: A Continuing SAGA. In: T. Gomi (Ed.): Evolutionary
Robotics. From Intelligent Robotics to Artificial Life. ER 2001, LNCS 2217, pp. 94–
109. Springer-Verlag Berlin Heidelberg 2001
References
221
Hazen, R.M., 2001. Selective adsorption of L- and D-amino acids on calcite: Implications
for biochemical homochirality. Proceedings of the National Academy of Sciences
98: 5487-5490.
Hebb, D.O., 1949. The organisation of behaviour. New York, Wiley.
Hebb, D.O., 1955. Drives and the conceptual nervous system. Psychological Review 62:
243-254.
Heino, M., Metz, J.A.J. and Kaitala, V., 1998. The enigma of frequency-dependent
selection. Trends Ecol. Evol. 13: 367-370.
Hengeveld, R. & Fedonkin, M.A., 2007. Bootstrapping the energy flow in the beginning of
life. Acta Biotheoretica 55: 181-226.
Hengeveld, R., 2007. Two approaches to the study of the origin of life. Acta Biotheoretica,
55: 97-131.
Herre, E.A., Knowlton, N., Mueller, V.G. & Rechner, S.A., 1999. The evolution of
mutualisms: exploring the paths between conflict and cooperation. - Trends Ecol.
Evol. 14: 49-53.
Hernández-Zapatha, E., Martinez-Balbuena, L. & Santamaria-Holek, I., 2009.
Thermodynamics and dynamics of the formation of spherical lipid vesicles. J. Biol.
Phys. 35: 297-308.
Heylighen, F., 1989a. Self-organisation, emergence and the architecture of complexity. In
‘Proceedings of the 1st European conference on System Science’, (AFCET, Paris)
pp. 23-32.
Heylighen, F., 1989b. Building a science of complexity. Proceedings of the 1988 Annual
Conterence of the Cybernetic Society (London).
Heylighen, F., 1990. Relational Closure: a mathematical concept for distinction-making
and complexity analysis. In: Cybernetics and Systems '90 (ed. R. Trappl). World
Science, Singapore, pp. 335-342.
Heylighen, F., 1991. Modeling Emergence. World Futures: the Journal of General
Evolution 31: 89–104.
Heylighen, F., 1995. (Meta)Systems as constraints on variation. A classification and
natural history of metasystem transitions. World Futures: The journal of general
evolution 45: 59-85.
Heylighen, F., 2000. Evolutionary transitions: how do levels of complexity emerge?
Complexity 6: 53-57.
Heylighen, F., 2006: "Mediator Evolution: a general scenario for the origin of dynamical
hierarchies", in: D. Aerts, B.D'Hooghe & N. Note (eds.) Worldviews, Science and
Us. (Singapore: World Scientific)
Heylighen, F., 2010 (in press). The self-organization of time and causality: steps towards
understanding the ultimate origin. Foundations of Science
Høgh-Jensen, H., 1998. Systems theory as a scientific approach towards organic
farming. Biological agriculture and horticulture 16: 37-52.
Holland, J.H., 1998. Emergence from chaos to order. Oxford University Press, Oxford.
Hooft ‘t, G., 1992. De bouwstenen van de schepping. Een zoektocht naar het
allerkleinste. Ooievaar Pockethouse Amsterdam.
222
References
Hooft ‘t, G., 2000. De fysica en het wonderbaarlijke. Nederlands Tijdschrift voor
Natuurkunde 2000: 316-319.
Hooke, R., 1665. Micrographia, or some physiological descriptions of minute bodies
made by magnifying glasses. London: J. Martyn and J. Allestry, 1665.
Horgan, J., 1995. From complexity to perplexity. Scientific American, 272: 74-79.
Hull, D.L., 1980. Individuality and selection. Annu. Rev. Ecol. Syst. 11: 311-332.
Hull, D.L., 1981, Units of evolution; a metaphysical essay. In U.J. Jensen and R. Harré,
eds. The philosophy of evolution. New York; St. Martins Press, pp. 23-44.
Ideker, T., Galitski, T., & Hood, L., 2000. A new approach to decoding life: systems
biology. Annual Review of Genomics and Human Genetics 2: 343-372.
Jagers op Akkerhuis, G.A.J.M. & van Straalen, N.M., 1999. Operators, the Lego–bricks of
nature: evolutionary transitions from fermions to neural networks. World Futures,
the Journal of General Evolution 53: 329-345.
Jagers op Akkerhuis, G.A.J.M & Damgaard, C.F., 1999. Using resource dominance to
explain and predict evolutionary success. Oikos 87: 609-614.
Jagers op Akkerhuis, G.A.J.M., 2001. Extrapolating a hierarchy of building block systems
towards future neural network organisms. Acta Biotheoretica 49: 171-189.
Jagers op Akkerhuis, G.A.J.M., 2001. In: Bloemmen, M. (Ed.): Essaybundel
Kennisdisseminatie: Technische intelligentie is volgend stadium in de evolutie.
Wageningen UR.
Jagers op Akkerhuis, G.A.J.M., 2008. Analysing hierarchy in the organisation of biological
and physical systems. Biological Reviews 83: 1-12.
Jagers op Akkerhuis, G.A.J.M., (2009). The operator hierarchy: evolution beyond genes.
Poster contribution. ECAL, 2009, Budapest, levels of selection workshop.
(http://www.sense.ecs.soton.ac.uk/levels-of-selection-workshop2009/abstracts/Jagers.pdf).
Jagers op Akkerhuis, G.A.J.M., (2010a, in press). Towards a hierarchical definition of life,
the organism, and death. Foundations of Science.
Jagers op Akkerhuis G.A.J.M., (2010b, in press). Explaining the origin of life is not
enough for a definition of life. Foundations of Science.
Jaros, G.G. & Cloete, A., 1987. Biomatrix, the web of life. World Futures 23: 203–224.
Jeuken, M., 1975. The biological and philosophical definitions of life. Acta Biotheoretica
24: 14-21.
Kaiser, D., 2001. Building a multicellular organism. Annual Review Genetics. 35: 103-123.
Kaku, M. & Thompson, J., 1987. Beyond Einstein: The cosmic quest for the theory of the
universe. Anchor Books, New York.
Kaku, M., 1994. Hyperspace. A scientific odyssey through parallel universes, time warps,
and the 10th dimension (Oxford University Press, Oxford)
Kauffman, S.A., 1986. Autocatalytic sets of proteins. Journal of Theoretical Biology 119:
1-24.
Kauffman, S.A., 1993. The origins of order: Self-organisation and selection in evolution.
(Oxford University Press, Oxford) (709 pp).
th
Keeton, W.T. & Gould, J.L, 1993. Biological science. 5 edition. W.W. Norton &
Company, Inc. NY.
References
223
Keilin, D., 1959. The Leeuwenhoek lecture: The problem of anabiosis or latent life:
History and current concept. Proceedings of the Royal Society of London, Series
B, Biological Sciences 150: 149-191.
Kelly, K., 1994. Out of control. The rise of neo-biological civilisation. A William Patric
Book. Addison Wesley Publishing Company, Reading.
Klijn, J.A., 1995. Hierarchical concepts in landscape ecology and its underlying
disciplines. Report 100, DLO Winand Staring Centre, Wageningen (The
Netherlands).
Knoll, M. & Ruska, E., 1932a. Das Elektronenmikroskop. (The electron microscope.) Z.
Physik 78: 318-339.
Knoll, M. & Ruska, E., 1932b. Beitrag zur geometrischen Elektronenoptik I und II.
(Contribution to geometrical electron optics.) Ann. Physik 12: 607-640 & 641-661.
Knoll, A.H., 1992. The early evolution of eukaryotes, a geological perspective. Science,
256: 622-627.
Koestler, A., 1967. The gost in the machine. London, Hutchinson & Co.
Koestler, A., 1978. Janus: a summing up. Hutchinson & Co. Ltd. London.
Kooijman, S.A.L.M., 1993. Dynamic energy budgets in biological systems. Cambridge
University Press, Cambridge, United Kingdom.
Korn, R.W., 2002. Biological hierarchies, their birth, death and evolution by natural
selection. Biology and philosophy 17: 199-221.
Korzeniewski, B., 2005. Confrontation of the cybernetic definition of a living individual with
the real world. Acta Biotheoretica 53: 1-28.
Koshland, D.E. Jr., 2002. The seven pillars of life. Science 295: 2215-2216.
Kot, M., Schaffer, W.M., Truty, G.L., Graser, D. J. & L. F. Olsen, 1988. Changing criteria
for imposing order. Ecol. Model. 43: 75-110.
Kragh, H., 1999: Videnskab og virkelighed. Aktuel Naturvidenskab; Århus, 1: 36-38.
Kruijff, H.A.M. de, 1991. Extrapolation through hierarchical levels. Comparative
Biochemistry and Physiology 100C: 291-299.
Kunin, V., 2000. A system of two polymerases. A model for the origin of life. Origins of
Life and Evolution of the Biosphere 30: 459-466.
Kurzweil, R, 1999. The age of the spiritual machines. When computers exceed human
intelligence. Viking, New York.
Labson, S., 1985. The ontology of science: an essay towards a complete description of
the universe. World Futures 21: 279-337.
Lane, N., 2005. Power, sex, suicide. Mitochondria and the meaning of life. Oxford
University Press, Oxford, pp. 354.
Laszlo, E., 1972. Introduction to systems philosophy: towards a new paradigm of
contemporary thought. Gordon and Breach, NY, London.
Laszlo, E., 1994. From GUTs to GETs: prospects for a Unified Evolution Theory. World
Futures 42: 233-239.
.
Laszlo, E., 1996. Evolution, the general theory Hampton Press, New Jersey.
Laszlo, E., 1996. The systems view of the world: a holistic vision for our time. Hampton
Press, Inc.
Lewontin, R.C., 1970. The units of selection. Rev. Ecol. Syst. 1: 1-18.
224
References
Lifson, S., 1987. Chemical selection, diversity, teleonomy and the second law of
thermodynamics. Reflections on Eigen’s theory of self-organisation of matter.
Biophysical Chemistry 26: 303-311.
Linowes, D.F., 1990. Speech delivered to the White House Conference on Libraries and
Information Services, Oct 1990. In: Dahlman, C. 1994. The third industrial
revolution: trends and implications for developing countries. Paper presented at
the Foro Nacional International Conference on the New International Order, 13–14
Apr, Rio de Janeiro, Brazil.
Mackie, G.O., Anderson, P.A.V. & Singla C.L., 1984. Apparent absence of gap junctions
in two classes of Cnidaria. The Biological Bulletin 167: 120-123.
Malthus, T.R., 1798. An essay on the principle of population. Oxford World's Classics
reprint.
Manasson, V.A., 2008. Are Particles Self-Organized Systems? arXiv:0803.3300v1
[physics.gen-ph]
Margulis, L, 1970. Origin of Eukaryotic Cells, Yale University Press, New Haven, CT.
Margulis, L., 1981. Symbiosis in cell evolution. New York: W.H. Freeman.
Margulis, L., 1998. The symbiontic planet. A new look at evolution. Weidenfeld &
Nicolson, London.
Martin, W. & Müller, M., 1999. The hydrogen hypothesis for the first eukaryote. Nature
392: 37-41.
Martin, W., Hoffmeister, M., Rotte, C. & Henze, K., 2001. An overview of endosymbiotic
models for the origins of eukaryotes, their ATP-producing organelles
(mytochondria and hydrogenosomes), and their heterotrophic lifestyle. Biol. Chem.
382: 1521-1539.
Martin, W., Russell, M.J., 2003. On the origins of cells: a hypothesis for the evolutionary
transitions from abiotic chemistry to chemoautotrophic prokaryotes, and from
prokaryotes to nucleate cells. Philosophical Transactions of the Royal Society of
London, series B-Biological Sciences 358: 59-83.
Maturana, H.R., 1969. The neurophysiology of cognition. In: Cognition: A multiple view (P
Garvin Ed.), Spartan Books, New York.
Maturana, H.R. & Varela, F.J., 1980. Autopoiesis and Cognition. The Realisation of the
Living. Dordrecht: D. Reidel, (also in Boston Studies in the Philosophy of Science,
42).
Maynard Smith, J.M., Szathmáry, E., 1995a. The Major Transitions in Evolution. Freeman
& Co., Oxford; now Oxford University Press.
Maynard Smith, J.M., Szathmáry, E., 1995b. The major evolutionary transitions. Nature
374: 227-231.
Maynard Smith, J.M., Szathmáry, E., 1999. The origins of life. Oxford University Press,
Oxford.
Mayr, E., 1982. The growth of biological thought: diversity, evolution, and inheritance.
Harvard University Press, Cambridge, Massachusetts, London, England.
Mazzarello, P., 1999. A unifying concept: the cell theory. Nature Cell Biology 1: E13-E15.
McCormick, D.A., 1995. The cerebrellar symphony. Nature 374: 412-413
References
225
McShea, D.W., 1991. Complexity and evolution: what everybody knows. Biol. Philos. 6:
303-324.
Mendeleev, D., 1869. ‘The Relation between the Properties and Atomic Weights of the
Elements’, Journal of the Russian Chemical Society, 1: 60-77 [Engl. transl.]
Mendeleev, D.I., 1871. Die periodische Gesetzmaessigkeit der chemischen Elemente.
Liebigs Annalen der Chemie. Suppl. 8(2): 133-229.
Michod, R.E., 1999. Darwinian dynamics. Evolutionary transitions in fitness and
individuality. - Princeton Univ Press, Princeton, NJ.
Miller, J.G., 1978. Living systems. New York: McGraw-Hill.
Mills, D.R., Peterson, R.I. & Spiegelman, S., 1967. An extracellular Darwinian experiment
with a selfduplicating nucleic acid molecule. Proceedings of the Natrional
Academy of Sciences of the USA 58: 217-224.
Morales, J., 1998. The definition of life. Psychozoan 1: (a strictly electronic journal) 1-39.
Mountcastle, V.B., 1975. An organizing principle for cerebral functioning. The unit module
and the distributed system. In: G.M. Edelman & V.B. Mountcastle, Eds. The
mindful brain. Cambridge, MA: MIT press.
Munson, J.R., York, R.C., 2003. Philosophy of biology. Encyclopedia Britannica.
Britannica 2003 ultimate reference suite.
Murre, J.M.J., Phaf, R.H., Wolters, G., 1992. CALM: Categorizing and learning module.
Neural Networks 5: 55-82.
Murré, J.M.J., Phaf, R.H., Wolters, G., 1989. CALM networks: a modular approach to
supervised and unsupervised learning. Pp. 649-656. Proc Int Joint conf Neural
Networks, Washington DC. New York IEEE Press.
Murre, J.M.J., 1992. Learning and categorisation in modular neural networks. HempelHempstead: Harvester Wheatsheaf (Hillsdale, NJ: Lawrence Erlbaum).
Naveh, Z. & Lieberman, A.S., 1994. Landscape ecology. Theory and application. 2e
edition. Springer-Verlag, New York.
Naveh, Z., 2000. What is holistic landscape ecology? A conceptual introduction.
Landscape and urban planning 50: 7-26.
Nederbragt, H., 1997. Hierarchical organisation of biological systems and the structure of
adaptation in evolution and tumorigenesis. J. Theor. Biol. 184: 149-156.
Newman, M.C., Jagoe C.H., 1996. Eds. Ecotoxicology, a hierarchical treatment. CRC
Press, Boca Raton.
Neumann, J. von & Burks, A.W., 1966. ‘Theory of Self-Reproducing Automata.’
University of Illinois Press. Champaign, IL, USA
Nicholson, B.J., 2003. Gap junctions – from cell to molecule. Journal of Cell Science 116:
4479-4481.
Occhialini, G.P.S. & Powel, C.F., 1947. Nuclear disintegrations produced by slow charged
particles of small mass. Nature 159: 186-190.
Odum, H.T., 1994. Ecological and General Systems: An Introduction to Systems Ecology.
University Press of Colorado, Niwot. 644 pp.
d
Odum, E.P, 1959. The fundamentals of ecology. 2 edition. Sounders, Philadelphia,
Pennsylvania.
226
References
Oliver, J.D. & Perry, R.S., 2006. Definitely life but not definitely. Origins of Life and
Evolution of the Biosphere 36: 515-521.
Oparin, A.I., 1957. The origin of life on earth. Academic press, New York.
Overton, E., I895. Uber die osmotischen Eigenschaften der lebenden Pflanzen und
Tierzelle. Viertjahr-schr. d. naturforsch. Ges. in Zurich, 40:159.
Pagels, H.R., 1985. Perfect symmetry: the search for the beginning of time. New York.
Simon and Schuster.
Panchin, Y.V., 2005. Evolution of gap junction proteins – the pannexin alternative.
Journal of Experimental Biology 208: 1415-1419.
Peracchia, C. & Benos D.J. (eds.), 2000. Gap Junctions: Molecular Basis of Cell
Communication in Health and Disease (Current Topics in Membranes, Volume 49)
by Academic press, New York.
Perl, M.L., 1987. The experimental view of particles and forces in 1987: a picture outline
of the talk. (Talk presented at the annual meeting of the American Association for
advancement of Science, Chicago, Illinois February 1987)
Plotkin, H.C., 1994. Darwin machines and the nature of knowledge. Penguin Books,
London.
Prigogine, I. & Stengers, I., 1984. Order out of chaos. Man’s new dialogue with nature.
Bantam, New York.
Pross, A., 2003. The driving force for life’s emergence: kinetic and thermodynamic
considerations. J. Theor. Biol. 220: 393-406.
Ponge, J.E., 2005. Emergent properties from organisms to ecosystems: towards a
realistic approach. Biological Reviews 80: 403-411.
Popa, R., 2003, Between necessity and probability. Searching for the definition and the
origin of life. Advances in Astrobiology and Biogeophysics. Springer.
Popper, K.R. 1972. Objective Knowledge: An Evolutionary Approach. London, Oxford
Univ. Press, 380 pp.
Popper, K.R. 1999. All Life Is Problem Solving. Routledge. 192 pp.
Poundstone, W., 1984. The Recursive Universe: Cosmic Complexity and the Limits of
Scientific Knowledge. New York, William Morrow.
Ray, T.S., 1991. Evolution and optimisation of digital organisms. In: Billingsley K. R.,
Derohanes E., H. Brown, III [eds.], Scientific Excellence in Supercomputing: The
IBM 1990 Contest Prize Papers, Athens, GA, 30602: The Baldwin Press, The
University of Georgia. pp. 489-531.
Rivera, M., Jain, R., Moore, J.E. & Lake, J.A., 1998. Genomic evidence for two
functionally distinct gene classes. Proceedings of the National Academy of
Sciences of the USA 95: 6239-6244.
Rosen, R., 1958. A relational theory of biological systems. Bulletin of Mathematical
Biophysics 20: 245-260.
Rosen, R., 1973. On the dynamical realisation of (M,R)-systems. Bulletin of Mathematical
Biophysics 35: 1-9.
Rosen, R., 1991. Life itself. A comprehensive inquiry into the nature, origin and
fabrication of life. Columbia University Press.
References
227
Ruska, E. 1986. The development of the electron microscope and of electron microscopy.
Nobel lecture, December 8, 1986.
Russell, B. 1960. An Outline of Philosophy. Meridian Books, Cleveland Ohio.
Rutherford, E. & Soddy F., 1902. The Cause and Nature of Radioactivity. From
Philosophical Magazine 4, 370-96 [as abridged and reprinted in Henry A. Boorse
& Lloyd Motz, The World of the Atom, Vol. 1 (New York: Basic Books, 1966.
Rutherford, E., 1911. The Scattering of the α and β Rays and the Structure of the Atom.
From the Proceedings of the Manchester Literary and Philosophical Society, IV,
55, pp. 18-20 Abstract of a paper read before the Society on March 7, 1911.
Ruiz-Mirazo, K., Pereto, J. & Moreno, A., 2004. A universal definition of life: autonomy
and open-ended evolution. Origins of Life and Evolution of the Biosphere 34: 323345.
Sagasti, F. & Acalde, G., 1999. Development cooperation in a fractured global order. An
arduous transition. International development research centre, Ottawa, Canada.
Schleiden, M.J., 1838. Beiträge zur Phytogenesis. Archiv für Anatomie, Physiologie und
wissenschaftliche Medicin 2: 137-176
Schneider, E. D., 1988. Thermodynamics, ecological succession, and natural selection: a
common thread. In: Entropy, information, and evolution, new perspectives on
physical and biological evolution. Eds.: B. H. Weber, D. J. Depew and J. D. Smith.
pp 108-138. Boston: MIT Press.
Schneider, E.D. & Kay, J.J., 1994. Life as a Manifestation of the Second Law of
Thermodynamics. Mathematical and Computer Modelling, Vol 19: 25-48.
Schreiber, S.J. & Gutierrez, A.P., 1998. A supply/demand perspective of species
invasions and coexistence - applications to biological control. Ecol. Model. 106:
27-45.
Schrödinger, E., 1944. What is life? Cambridge University Press.
Schultz, A.M., 1967. The ecosystem as a conceptual tool in the management of natural
resources. In: ‘Natural resources: quality and quantity’, pp 139-161. Univ of
California Press, Berkeley.
Schwann, T., 1839. Mikroskopische Untersuchungen über die übereinstimmung in der
Struktur und dem Wachstum der Tiere un Pflanzen. Sander’schen Buchhandlung,
Berlin.
Schwartz, R.M. & Dayhoff M.O., 1978. Origins of prokaryotes, eukaryotes, mitochondria
and chloroplasts. Science 199: 395-403.
Searcy, D.G., 2003. Metabolic integration during the evolutionary origin of mitochondria.
Cell Research, 13: 229-238.
Shannon, C.E., 1948. A mathematical theory of communication. Bell System Technical
Journal, vol. 27: 379-423 and 623-656.
Sheehan, P.J., 1984. Effects on community and ecosystem structure and dynamics. In:
Effects of pollutants at the ecosystem level. Eds. Sheehan P.J., Miller D.R. et al.
SCOPE , John Wiley & Sons Ltd.
Silvestre, D.A.M.M. & Fontanari, J.F. 2008. The information capacity of hypercycles.
Journal of Theoretical Biology 254: 804-806.
228
References
Simon, H.A., 1962. The architecture of complexity. Procedings of the American Society
for the Philosophy of Science 106: 467-482.
Simon, H. A., 1973. The organisation of complex systems. In: Hierarchy theory (ed. H. H.
Pattee), pp. 3–27. George Braziller, New York.
Sims, K., 1994. ‘Evolving 3D Morphology and Behavior by Competition’ IN: Artificial Life
IV Proceedings, ed. by Brooks & Maes, MIT Press, pp. 28-39.
Wilson, D.S. & Wilson, E.O., 2007. Rethinking the theoretical foundation of socio-biology.
The Quarterly Review of Biology 82: 327-348.
Sole, R.V., Manrubia, S.C., Benton, M., Kauffman, S. & Bak, P., 1999. Criticality and
scaling in evolutionary ecology. - Trends Ecol. Evol. 14: 156-160.
Spiegelman, S., 1971. An approach to the experimental analysis of precellular evolution.
Quarterly Reviews of Biophysics 4: 213-253.
Spier, F., 1996. The Structure of Big History: From the Big Bang until Today, Amsterdam
University Press, Amsterdam.
Steen, W.J. van der, 1997. Limitations of general concepts: a comment on Emmeche’s
definition of ‘life’. Ultimate Reality and Meaning 20: 317-320.
Stokes Brown C., 2007. Big History: From the Big Bang to the Present. New York : New
Press.
Straalen, N.M. van & Roelofs, D., 2006. An Introduction to Ecological Genomics. Oxford
University Press, Oxford.
Swenson, R., 1989. Emergent attractors and the law of maximum entropy production:
Foundations to a theory of general evolution. Systems research 6: 187-197.
Szentagothai J., 1975. The ‘module concept’ in the cerebral cortex architecture. Brain res.
95: 475-496
Teilhard de Chardin, P. 1959. The Phenomenon of Man. Harper, New York. (Original:
Editions du Seuil I, Paris, 1955).
Teilhard de Chardin, P, 1966. Man's place in nature. The human zoological group.
Collins, London. (Original: Editions du Seuil VIII, Paris, 1949).
Teilhard de Chardin, P., 1969. The future of man. Collins, London. (Original: Editions du
Seuil V, Paris,1946)
Tinbergen, N., 1946. Inleiding tot de diersociologie (Introduction to animal sociology). J.
Noorduijn en Zoon, Gorinchem.
Townsend, C.R., Begon, M. & Harper, J.L., 2008. Essentials of ecology. Third edition.
Blackwell publishing.
Turchin, V.E., 1977. The phenomenon of science, a cybernetic approach to human
evolution. Colombia University Press
Turchin, V., 1977. The Phenomenon of Science. New York: Columbia University Press.
Turchin, V., 1995. A dialogue on metasystem transitions. In: Heylighen, F., Joslyn, C.,
and Turchin, V. (eds.): The Quantum of Evolution. Toward a theory of metasystem
transitions.(Gordon and Breach Science Publishers, New York) (special issue of
‘World Futures: the journal of general evolution, volume 45, 1995).
Ulrich, L.P., 1972. The concept of man in Teilhard de Chardin. PhD thesis, University of
Toronto, Canada.
References
229
Ulanowicz, R.E., 2009. The dual nature of ecosystem dynamics. Ecological modelling
220: 1886-1892.
Valen, L., 1973. A new evolutionary law. Evolutionary Theory 1: 1-30.
Varela, F.J., 1976. Not one, not two. The CoEvolution Quarterly, 11: 62-67.
Varela, E.J., 1979. Principles of biological autonomy. In: Dr. G. Klir (Ed.). The north
Holland series in general systems research.
Varela, F.J., 1984. The creative circle: Sketches on the natural history of circularity. In P.
Watzlawick (Ed.), The invented reality: How do we know what we believe we
know? Contributions to constructivism (pp. 309-323). New York: Norton.
Virchow, 1855. Die Cellular-Pathologie. Archiv der pathologischen Anatomie, Physiologie
und klinischen Medizin. 1: 3-39.
Valentine, J.W., 2003. Architectures of biological complexity. Integr. Comp. Biol. 43: 99103.
Verhulst, P-F., 1838. "Notice sur la loi que la population poursuit dans son
accroissement" (PDF). Correspondance mathématique et physique 10: 113–121.
Vidal, C. 2009. Metaphilosophical Criteria for Worldview Comparison. ECCO Working
Paper (http://tinyurl.com/Vidal-MetaCriteria).
Wächtershäuser, G., 2000. Origin of life: Life as We Don't Know It. Science 289 (5483):
1307 – 1308.
Wagner, G.P., 1996. Does evolutionary plasticity evolve? Evolution 50: 1008-1023.
Wagner, G.P. & Altenberg, L., 1996. Complex adaptations and the evolution of
evolvability. Evolution 50: 967-976.
Walker, J., 2009. Atoms, a short history of the knowledge of the atom.
www.nobeliefs.com/atom.htm
Wallace, A.R., 1855. On the law which has regulated the introduction of new species. Ann. Mag. Nat. Hist. ser. 2 16: 184-196.
Warwick, K., 1997. The march of the machines. Why the new race of robots will rule the
world. Century Books Ltd., London.
Webster, J.R., 1979. Hierarchical organisation of ecosystems. In: Theoretical systems
ecology (ed. E. Halfton), pp. 119-129. Academic Press Inc.
Whitehead, A.N., 1925. Science and the Modern World. (1997 paperback, Free Press,
Simon & Schuster.)
Whitehead, A.N., 1929. Process and Reality: An Essay in Cosmology. (1979 corrected
edition, edited by David Ray Griffin and Donald W. Sherburne, Free Press)
Weiner, J., 1990. Asymmetric competition in plant populations. - Trends Ecol. Evol. 5:
360-364.
Weinberg, S., 1977. The first three minutes. A modern view of the origins of the universe.
Basic Books, New York.
Weinberg, S., 1992. Dreams of a final theory. Pantheon books.
Weiss, P.A., 1969. The living system: Determinism stratified. In: Beyond reductionism
(eds. A. Koestler &J. R. Smythies), pp. 3-55. Hutchinson. London.
Weiss, P.A., 1971. Hierarchically organized systems in theory and practice. Hafner Publ.
Comp., New York.
Wilczeck, F., 1998. The standard model transcended. Nature 394: 13-15.
230
References
Wilkins, J.S., 1998. What’s in a meme? Reflections from the perspective of the history
and philosophy of evolutionary biology. J. Memetics 2: 1-29.
Willensdorfer, M., 2008. On the evolution of differentiated multicellularity. arXiv,
0801.2610v1
Wilson, A., 1969. Closure, entity and level. In: Hierarchical structures. (L.L. Whyte, A.G.
Wilson and D. Wilson, eds.) pp 54-55. Am. Elsevier, New York.
Wilson, E.B., 1892. The cell in development and inheritance. New York: MacMillan.
Wolpert, L.1995. Evolution of the cell theory. Philosophical transactions of the royal
society of London, Series B: Biological Sciences 349: 227-233.
Wilczek, F., 1998. The standard model transcended. Nature 394: 13-15.
Witten, E., 1988. In: Davies, P.C.W. & Brown, J.R. (eds.): Superstrings: A theory of
everything? Cambridge University Press.
Witting, L., 1997. A general theory of evolution by means of selection by density
dependent competitive interactions. - Peregrine Publ., Arhus.
Yukawa, H., 1935. ‘On the Interaction of Elementary Particles. I.’ Proc. Phys.-Math. Soc.
Jpn. 17: 48.
Zwart H., (2010). Pervasive Science: Challenges of contemporary technosciences for
governance and self-management. In: M.E.A. Goodwin et al (eds.) /Dimensions of
Technology /Regulation.Wolf Legal Publishers, 189-202.
Summary
“To confine our attention to terrestrial matters would be to limit the human spirit.”
(Stephen Hawking)
232
Summary
It has been estimated that more discoveries have been made in the past 30
years than in the rest of human history. This explosion of knowledge has caused
a crisis in science that is accompanied by overspecialisation and
compartmentalisation. Scientists face the challenge of overcoming this crisis by
constructing comprehensive and integrating concepts. This thesis presents a
new method for creating bridges between disciplines and fighting the knowledge
explosion crisis.
In human history various theories have been developed that aim at creating
coherence in the phenomena observed in the world. An ancient approach is that
of the Ouroboros, the snake swallowing its tail. This symbol represents the
continuity and eternity of life and the world. Later, in his Scala Naturae, the
Greek philosopher and naturalist Aristotle ranked natural phenomena by
decreasing perfection, from spiritual and divine beings to man, animals, plants
and finally rocks and formless matter. Aristotle’s vision is also referred to as the
Great Chain of Being. The Great Chain markedly influenced medieval
Christianity, and, in turn, this chain influenced several well-known philosophers
such as Descartes, Spinoza and Leibniz.
A current, comprehensive approach to the organisation of nature is Big History.
Big History ranks all systems and processes by their occurrence in cosmic
history. The approach rests on the scientific theory that the early universe was
as small as a fundamental particle and obtained its present size after an
explosive expansion: the Big Bang. Although the question about the universe’s
origin remains unanswered, the Big Bang theory is firmly supported by modern
particle physics and cosmological observations.
Sometimes old structuring attempts are given a new interpretation, such as the
Ouroboros in Figure 1. The left half of the figure shows how science has gained
access to increasingly small systems, from large animals and plants, to
individual cells, molecules, atoms, nuclear particles and finally quarks. The right
half of the figure shows the scope of scientific research broadening towards
increasingly large systems, from ecosystems, through planets, to solar systems,
galaxies and finally the entire universe. When presented in this way, specific
properties of the universe at large, such as the background radiation and its
expansion in all directions, are connected with the universe’s initial state. As the
Ouroboros picture suggests, where the enormous and the miniscule meet, the
search for organisation in nature seems to swallow its tail.
Summary
233
Figure 1. This figure shows the historical expansion of scientific domains in the direction
of the smallest systems and in the direction of the largest systems. GUT = grand unifying
theory, Z and W indicate fundamental force carrying particles that are responsible for the
weak nuclear force, DM = dark matter, cm = centimetre.
Evolution beyond Darwin
Despite all efforts aimed at unification in science, a general theoretical
framework connecting the many fields of science has yet to be found. Many
regard the lack of such a framework as a major deficit because it reduces our
abilities to counteract current centrifugal trends in the knowledge explosion. This
thesis aims to tackle this problem by systematically exploring unifying principles
in science that may allow us to connect various scientific domains such as
particle physics and evolution.
Darwin’s theory of evolution offers a good starting point to explain how this
thesis has approached the goal of developing unifying concepts. Darwin stated
234
Summary
that selection acts against offspring that are less well adapted to their
environment. A white butterfly living in an ecosystem full of dark volcanic rocks is
not well adapted to its environment because predatory birds can easily spot it.
Consequently, it will suffer more predation than its darker relatives. As the genes
of the dark parents code for dark offspring, the percentage of dark butterflies will
increase with every generation. This example shows that the theory of evolution
uses two processes. Firstly, the parents produce offspring with unique
combinations of heritable properties and then the environment selects against
less well-performing individuals.
Butterflies and other organisms are not the only things that evolve. Particles
evolve, too. Of course, the evolution of particles does not involve heritable
variation. However, if particles collide, the product may be a new particle or a
range of new particles. Just like reproduction, such processes produce new
variation, and the particles that are produced also experience selection. For
example, the hot gaseous environment of a star continuously produces various
atomic “species”, while only a few of these are stable in the hot environment. We
may regard this as a form of evolution where the process of diversification is
based on production instead of on reproduction.
Productive and reproductive evolution processes became active at different
moments in cosmic history. The early universe had only physical particles and
only production could create diversity that selection could act on. Much later, the
emergence of organisms introduced the possibility of reproductive creation of
diversity.
The difference between the productive and the reproductive production of
diversity coincides with the conceptual separation of physics and biology. To
close this gap, this thesis suggests using a general definition of evolution based
on diversification (the production of variation) and selection, and the modification
of these aspects by circumstances. In relation to given contexts, diversification
can relate to the chemical production of variation, to organic reproduction and, in
the future, to the technical construction of offspring. Similarly, selection can
change in relation to the system types that are considered. This approach allows
the evolution concept to be applied to all operators in the universe, from
fundamental particles through living organisms to future robotic life forms.
Levels of complexity
The difference between evolution based on production and reproduction is,
however, not the only reason why the fields of physics and biology seem to
resist integration. There is another barrier. In physics, researchers usually have
a clear picture in mind when referring to levels of complexity, such as nuclear
Summary
235
particles, atoms and molecules. The formation of these particles is generally
attributed to condensation reactions and not to evolution. In biology, Darwin’s
theory of evolution is broadly accepted, but this theory ignores he complexity
levels of organisms.
Darwin knew, of course, that bacteria, protozoa, plants and fungi, and animals
existed. But he did not need the differences in complexity between these groups
to define what was later coined the “struggle for life” and “survival of the fittest”.
Although genetics has contributed much to the understanding of evolutionary
mechanisms, its development has not helped much with understanding the
complexity levels of organisms, with the exception of the difference between
organisms without or with mitochondrial genes.
That it pays no attention to structural complexity represents a serious flaw in
Darwin’s theory. Because of this flaw, the theory of evolution cannot answer an
apparently simple question such as “Is it possible to predict what the next step in
the evolution process will be?” It also cannot answer a question such as “Are
endosymbiontic/eukaryotic cells the only logical successors of bacteria?” To
answer these questions requires a comprehensive theory that accounts for
structural complexity and its levels.
Dimensions for hierarchy
Levels in the organisation of nature are frequently represented as a sequence of
increasingly complex system types. The literature contains many examples of
such sequences. Generally, these sequences exhibit marked similarity to the
following example: atom, molecule, organelle, cell, tissue, organ, organism,
population, community, ecosystem, planet, solar system, galaxy and universe
(Fig. 2). What is very interesting about these kinds of hierarchies is that, at face
value, they seem useful tools for organising complexity, whereas on closer
consideration it is easy to show that they are actually based on an inconsequent
use of hierarchy rules.
236
Summary
Figure 2. A typical example of the much
used ranking of complexity generally
indicated as ‘ecological hierarchy’
To solve problems with existing approaches, Figure 3 introduces a new vision to
analyse hierarchy in nature. It is based on three fundamental directions: (1) the
hierarchy from a particle to a system of interacting particles, (2) the hierarchy
ranking particles of different complexity, and (3) the hierarchy in the interior of a
particle. This thesis proposes that these three directions represent three
dimensions for hierarchy.
Summary
237
Figure 3. The figure represents the three fundamental dimensions for hierarchy in the
organisation of nature Upward pointing arrows indicate increasing complexity in systems
of interacting particles. Arrows pointing to the right indicate how lower level organisms
create higher level organisms. Downward pointing arrows indicate increasing complexity
in the internal organisation of an organism.
The idea that there are three dimensions for hierarchy in nature is illustrated in
Figure 3. For simplicity reasons, the figure is deliberately limited to prokaryotic
cells and eukaryotic cells. Based on the same principle, the figure can be
extended in the direction of fundamental particles and in the direction of more
complex organisms. The figure shows that it is important to use dimensions in a
systemic manner. It is, in principle, not allowed to switch between dimensions
when analysing hierarchy. In relation to the latter rule, Figure 4 shows what an
ecological ranking looks like that respects the three dimensions for hierarchy.
The left column of this figure shows that local areas in the universe display an
ongoing development and differentiation from vague aggregations of matter, to
galaxies, solar systems, planets and ecosystems. On the opposite side, the right
column illustrates how hierarchy can be analysed in the interior of a particle. The
column in the centre shows that specific forms of aggregation or the formation of
internal structures can give rise to more complex particles.
238
Summary
Figure 4. Ranking of systems that accounts for three dimensions for ecological hierarchy Towards a strict ranking
The particle hierarchy in the central column of Figure 4 forms the core of the
theory discussed in this thesis. In this hierarchy, every step results from a selforganisation process. This self-organisation automatically causes a new cyclic
process (a functional aspect) and a new circular shape (a structural aspect). A
concept generally used in system science to indicate cyclic interactions is
‘closure’. As the result of the functional and structural closure, the interactions
form a new unit.
The theory in this thesis now focuses on situations in nature in which a next
particle type and associated closures are realised as the first possibility nature
allows. This logic was named the principle of ‘first-next possible closure’. The
use of first-next possible closure allows the creation of a strict hierarchy that is
based on self-organisation steps, a property that is fundamental to the theory
presented in this thesis. Whereas most theories apply a rather loose and
intuitive classification of stages or levels of complexity, our objective is to
develop a systematic version that explains why the next level was both possible
and inevitable.
Summary
239
The definition of first-next possible closure allowed the identification of all
systems in the universe that comply with this demand. These systems form a
highly special and strictly hierarchical subset of all the systems in the universe.
This subset includes the quarks and hadrons, the atoms, the molecules, the
prokaryotic cells, the eukaryotic cells, the multicellulars and the animals.
Searching for a logical name for all the systems in this subset, I realised that,
due to their closures, these systems can be considered as individual operational
units. This was the reason why I named them ‘operators’. Thus, by first-next
possible closure, an operator type at one particular level of complexity
transforms to an operator type at the next level.
The introduction of the operator concept can be regarded as a major tool for
bringing physics and biology closer by bridging the gap between abiotic and
living systems; the operator concept shows that both spheres of existence are
ultimately guided by the same organising principles such that, at a high level of
abstraction, both physics and biology are studying the same kind of systems: the
operators. The word operator has various meanings in other contexts, such as
mathematics and the operation of machinery. In the context of the theory
proposed in this thesis, the concept always refers to the system types that are
related to each other by first-next possible closures.
The dark line in Figure 5 shows how subsequent first-next possible closures
connect all the operators. This hierarchical organisation has been named the
“operator hierarchy” and represents an important result of the present study. In
comparison to other approaches to hierarchy in nature, the operator hierarchy
possesses the special property that it is strict; that is, it is impossible to delete a
single operator from the scheme or add one to it without destroying the
scheme’s entire logic. Another aspect that is unique to the operator hierarchy is
that it introduces a new name for animals: memons. This adaptation is needed
to distinguish between the organisms that do and the organisms that do not
show neural network structure. Accordingly, a human being is an organism and
a memon, but a plant is only an organism. The concept of the memon is also
more practical than the concept of an animal when it comes to classifying future
intelligent technical beings.
240
Summary
Figure 5. The evolution of the operators
The dark line shows the historical pathway by which first-next possible closures created
the operators. The gray columns indicate systems resulting from first-next possible
closure that are not operators. Explanation of abbreviations: SAE ('Structural Auto
Evolution’) = the property of an operator to autonomously evolve the structure that carries
its information. SCI (‘Structural Copying of Information’) = the property of an operator to
autonomously copy its information (genes, learned knowledge) by simply copying part of
its structure. HMI (Hypercycle Mediating interface) = a closure creates an interface that
mediates the functioning of the hypercycle. Multi-state = operator showing closure
between multiple units of exactly one lower closure level. Hypercycle = closure based on
emergent, second-order recurrent interactions. Interface = closure creating an emergent
limit to an operator. CALM (Categorizing And Learning Module) = a minimum neural
memory.
A new natural order that connects scientific disciplines
The goal of this thesis was to systematically explore unifying principles in
science that may allow us to connect various scientific domains such as particle
physics and evolution. How does the operator hierarchy contribute to this goal?
Summary
241
As an answer to this question, I have summarised the ways in which the
operator hierarchy can be applied to various scientific integration topics:
 The operator hierarchy allows ecological hierarchies to be improved by taking
into account the following three dimensions for hierarchy: (1) the hierarchy of
the operators, (2) the internal organisation of the operators, and (3) the
interactions between operators, giving rise to hierarchy in large systems of
interacting operators. Uncontrolled mixing of these three aspects disturbs the
logic of any hierarchical ranking.
 Conventional definitions of Darwinian evolution are based on heredity,
reproductive diversification, and selection. The operator hierarchy now adds
levels of complexity to this list of aspects that are relevant to evolution.
 The operator hierarchy offers a periodic table to which other periodic tables
can be linked, such as the periodic table of the elements, the standard model,
the eightfold way of the hadrons, and the tree of life.
 The operator theory offers a completely new viewpoint for defining life, the
organism and death. Life was defined as matter with the operator structure
and an equal or even higher complexity than that of the cellular operator. In a
strict sense, an organism can be defined as an operator that agrees with the
definition of life. Death occurs when irreversible deterioration causes the loss
of the level of closure that is typical for the organism.
 If the operator hypothesis correctly predicts that the possibilities for
constructing all operators are guided by closure as a fundamental rule for
organising matter, the operator hierarchy would be valid in the entire
universe. If this is the case, extraterrestrial life will show the same
fundamental closures as life on Earth. Of course, the actual size, shape,
physiology, molecular construction, colour, etc. may differ.
 First-next possible closures may form and disintegrate. Given the existence
of a long sequence of operators, closure must represent a thermodynamically
favourable situation. Consequently, the arrow of complexity moves towards
complex operators, unless ambient conditions block the process.
 The operator hierarchy offers a framework that, possibly for the first time in
scientific history, allows essential construction-properties of future operators
to be predicted. The operator theory predicts that technical memons are the
next step in operator evolution. The use of the operator hierarchy gives much
more weight to the resulting predictions than can be given to extrapolations
based on existing trends in technological development. In addition, the
operator hierarchy opens up possibilities to look much further ahead.
 The operator hierarchy offers a structured scheme for discussing the things
between heaven and earth that are known to science and to speculate about
the possibilities for things still undiscovered.
 Because a memic architecture represents just another operator in the
sequence of the operator hierarchy, there are no scientific reasons for
considering human beings as some kind of final stage in evolution or as a
crown on a reputed creation.
242

Summary
The operator theory indicates that it may be practical to introduce a novel
interpretation of the meme concept. A meme currently stands for a cultural
replicator without any structural parallel with the gene, which stands for a
molecular replicator. By creating a meme concept that refers to the neural
network structure of a memon, the analogy can be improved. A structural
meme would take the form of a code-string coding for the number of neurons
in a network, the connections of these neurons and the inhibitory or excitatory
strengths of the connections. The new meme could actually code for all the
knowledge in a given neural network and would reintroduce Lamarckian
inheritance.
The operator theory: a new approach?
The operator theory is an innovative approach, but it is not an entirely new
concept. Around 1950, the French palaeontologist, philosopher and Jesuit Pierre
Teilhard de Chardin was working in the same direction. His attempt aimed at
integrating system theory with Christianity. In his work Teilhard de Chardin
distinguished the systems that are ‘formed’ and ‘centered’ from all the other
systems that do not show these properties. He was convinced that only the
formed and centred systems were relevant in creating an evolutionary hierarchy.
Teilhard de Chardin’s work is not generally accepted in the scientific community.
One reason for this may be that several important organising principles that
could have supported his approach still had yet to be discovered. Another
reason may be Teilhard de Chardin’s prediction of the unification of man and
God, called Point Omega, as a final level in the hierarchy.
The concepts of structural and functional closure and the use of operators in the
present approach show connections with the work of Teilhard de Chardin. The
operator theory extends the approach, however, with modern self-organisation
theory, with the concept of first-next possible closure and with the recognition of
the secondary structure of the operator hierarchy. In contrast to Teilhard de
Chardin’s work, the operator hierarchy strongly suggests that the evolution of the
operators has no end.
A detailed approach describing the evolution of particles and organisms
comparable to the operator hierarchy was not found in the literature. Because
this viewpoint is new, many of its aspects are still open to further research. I
consider it especially interesting and rewarding to increase our understanding of
the rules defining the operator hierarchy’s internal structure (Fig.5). Now that the
operator hierarchy offers a fundament for understanding the evolution of the
operators, the next challenging goal is to predict future operators as accurately
and as far ahead as possible.
Samenvatting
Samenvatting
“When I'm working on a problem, I never think about beauty.
I think only how to solve the problem.
But when I have finished, if the solution is not beautiful, I know it is wrong.”
(Buckminster Fuller)
243
244
Samenvatting
De laatste dertig jaar zijn meer ontdekkingen gedaan dan in alle eeuwen
daarvoor. Deze kennisexplosie heeft geleid tot een crisis in de wetenschap. De
afstand tussen verschillende disciplines neemt toe en het onderlinge begrip
neemt af omdat wetenschappers zich steeds meer zijn gaan specialiseren. Dit
proefschrift biedt een nieuw raamwerk dat samenhang creëert tussen
wetenschapsdisciplines en zo deze kenniscrisis te lijf kan gaan.
Door de eeuwen heen zijn verschillende theorieën ontwikkeld die zich richten op
het ordenen van fenomenen in de ons omringende wereld. Een voorbeeld uit de
verre oudheid is de Ouroboros, de slang die zijn eigen staart verzwelgt. Deze
slang was het symbool van de continuïteit en de eeuwigheid van het leven en de
wereld. Later, in zijn Scala Naturae stelde de Griekse filosoof Aristoteles een
indeling voor van de natuur op basis van afnemende perfectie, namelijk de
heilige en bovennatuurlijke zaken, mensen, dieren, planten en levenloze dingen.
Deze theorie speelt ook een rol in het middeleeuwse christendom en is terug te
vinden in het werk van bekende filosofen uit de gouden eeuw, zoals Descartes,
Spinoza en Leibniz.
Een recent voorbeeld van een ordening van de natuur is Big History. Big History
streeft naar een omvattende weergave van de ontwikkelingen in het heelal sinds
het allereerste begin. Het gaat ervan uit dat het vroege heelal zo klein was als
een fundamenteel deeltje en dat het tot zijn huidige grootte is uitgedijd na een
explosieve beginfase: de oerknal. Zowel de moderne deeltjesfysica als de
kosmologie leveren onderbouwing aan de oerknaltheorie.
Soms wordt aan oude symbolen van ordening een nieuwe inhoud gegeven,
bijvoorbeeld aan de Ouroboros in Figuur 1. Hier is aan de ene kant
weergegeven hoe de wetenschap inzicht heeft gekregen in steeds kleinere
systemen (van dieren, naar cellen, naar moleculen, atomen en quarks) terwijl ze
aan de andere kant steeds grotere systemen is gaan onderzoeken (van
ecosystemen, via planeten naar sterrenstelsels en uiteindelijk het hele heelal).
Deze manier van presentatie verbindt enkele bijzondere eigenschappen van het
heelal in zijn geheel, zoals de reststraling van de oerknal en de uitdijing in alle
richtingen, met de begintoestand van een mini-heelal, die de oorsprong was van
deze eigenschappen. Hierdoor lijkt het of het onderzoek naar de kosmische
ordening in zijn eigen staart bijt zoals bij de Ouroboros.
Samenvatting
245
Figuur 1: Historische uitbreiding van de reikwijdte van wetenschapsvelden in de richting
van steeds kleinere en steeds grotere systemen. GUT = grote unificerende theorie, DM =
donkere materie, Z en W staan voor twee fundamentele deeltjes die verantwoordelijk zijn
voor de zwakke kernkracht, cm = centimeter.
Darwin als startpunt
Ondanks al deze pogingen tot ordening is het nog niet gelukt om een algemeen
geaccepteerd raamwerk te ontwikkelen dat alle wetenschappelijke disciplines
met elkaar verbindt. Dit wordt vooral in de systeemkunde en de filosofie als een
gemis ervaren. Dit proefschrift beschrijft een nieuw ordenend principe met als
doel daarmee bruggen te kunnen bouwen tussen de verschillende wetenschapsvelden.
Bij het realiseren van dit doel is de evolutietheorie van Darwin als startpunt
genomen. Darwin stelt dat de natuur selecteert op nakomelingen die zijn
aangepast aan hun omgeving. Een witte vlinder die leeft in een omgeving met
246
Samenvatting
veel donkere rotsen, is slecht aangepast en vogels zullen hem snel herkennen.
De witte vlinder wordt eerder opgegeten dan zijn donkere soortgenoten. Op
deze wijze selecteert een donkere omgeving op donkere vlinders. Deze zullen
vervolgens via hun genen de donkere kleur doorgeven aan hun jongen en
generatie na generatie zal het percentage donkere individuen toenemen. Dit
voorbeeld laat zien dat de evolutie gebruik maakt van twee processen. Eerst
produceren ouders nakomelingen met een unieke mix van overerfbare
eigenschappen en vervolgens selecteert de omgeving in het nadeel van de
slechtst presterende individuen.
Vlinders en andere organismen zijn niet de enige zaken die evolueren. Ook
deeltjes evolueren. Natuurlijk is er bij deeltjes geen sprake van erfelijke variatie.
Maar als deeltjes samensmelten en soms weer opsplitsen zijn niet alle ontstane
deeltjes gelijk aan de oorspronkelijke deeltjes. Er ontstaan nieuwe varianten. En
ook deeltjes staan bloot aan selectie. Bijvoorbeeld in de hete gassen van een
ster worden voortdurend verschillende atoomkernen geproduceerd, terwijl
slechts sommige typen atoomkernen stabiel zijn in dit hete milieu. Ook hier is
sprake van evolutie, maar hierbij is de variatie in gevormde deeltjes het resultaat
van productie en niet van reproductie zoals bij de vlinders.
Deze twee vormen van evolutie hebben op verschillende momenten hun intrede
gedaan in het heelal. Eerst waren er alleen fysische deeltjes, zoals atomen en
moleculen en was productieve evolutie de enige mogelijkheid. Pas veel later
werd reproductieve evolutie mogelijk, namelijk nadat organismen met erfelijke
eigenschappen waren ontstaan.
Het verschil tussen evolutie op basis van productie en reproductie valt samen
met de scheiding tussen de fysica en de biologie. Om deze kloof te dichten, gaat
dit proefschrift uit van een algemene definitie van evolutie die de productie van
variatie combineert met selectie. De productie van variatie is breed toepasbaar
en kan opeenvolgend het resultaat zijn van chemische productie, van
biologische reproductie en zelfs, in de toekomst, van technische constructie van
nakomelingen. Het begrip selectie is eveneens bruikbaar in zowel de fysica als
in de biologie.
Niveaus in complexiteit
Het verschil tussen productieve en reproductieve evolutie is niet de enige
hinderpaal die in de weg staat bij integratie van fysische en biologische
wetenschapsvelden. Er is nog een barrière. In de fysica heeft men een helder
beeld van niveaus van complexiteit, zoals kerndeeltjes, atomen en moleculen.
Voor het ontstaan van deze niveaus denkt men gewoonlijk niet aan een
evolutieproces. In de biologie is de evolutietheorie van Darwin algemeen
Samenvatting
247
geaccepteerd, maar deze besteedt echter geen aandacht aan de complexiteit
van organismen.
Darwin kende zeer zeker de verschillen tussen bacteriën, protozoën,
meercelligen zoals planten en schimmels, en dieren. Maar hij had de verschillen
in complexiteit tussen deze groepen niet nodig om het mechanisme van
‘struggle for life’ en ‘survival of the fittest’ te bedenken. Diepgaande genetische
experimenten om de evolutietheorie van Darwin te onderbouwen hebben
evenmin bijgedragen aan een beter inzicht in niveaus van structurele
complexiteit.
Dit nu blijkt een tekortkoming van Darwin’s theorie. De evolutietheorie kan
namelijk in zijn huidige vorm geen antwoord geven op de simpele vraag
“Kunnen we de volgende stap in de evolutie voorspellen?” Ook andere vragen
zoals: “Zijn in de evolutie eukaryote cellen de enige logische opvolgers van
bacteriën, of zijn er andere mogelijkheden?” kunnen niet worden beantwoord.
Voor het oplossen van dergelijke fundamentele problemen is een theorie nodig
die inzicht geeft in het ontstaan van verschillende niveaus van organisatie in de
natuur.
Dimensies voor hiërarchie
Het ontstaan van organisatieniveaus wordt vaak weergegeven als een
hiërarchische volgorde van systemen. De literatuur geeft hiervan talrijke
voorbeelden. De gepresenteerde opeenvolgingen van systemen zijn op het
eerste gezicht heel logisch, zoals atoom, molecuul, organel, cel, weefsel,
orgaan, organisme, populatie, gemeenschap, ecosysteem, planeet,
zonnestelsel, sterrenstelsel, heelal ( figuur 2). Een dergelijke benadering van de
hiërarchie in de natuur heeft breed opgang gedaan in de exacte
wetenschappen. Echter, bij nadere bestudering blijkt deze hiërarchie niet
consequent.
In Figuur 3 wordt een nieuwe benadering geïntroduceerd voor het analyseren
van hiërarchie in de natuur op basis van drie fundamentele richtingen. Ten
eerste bestaat er een hiërarchie die loopt van individuele deeltjes naar groepen
van deeltjes. Ten tweede bestaat er een hiërarchie die loopt van fysische
deeltjes naar organismen. Ten derde bestaat er een hiërarchie die de interne
organisatie van de deeltjes ordent. Dit proefschrift stelt de term dimensies voor
voor deze drie manieren van ordening.
248
Samenvatting
Figuur 2. Een typische weergave van de veel
gebruikte ordening van complexiteit die meestal
wordt aangeduid als de ecologische hiërarchie.
In het voorbeeld van Figuur 3 staan alleen bacteriën en eukaryote cellen, maar
het is niet moeilijk om dit schema uit te breiden in de richting van fysische
deeltjes en in de richting van meer complexe organismen. Het voorbeeld in
figuur 3 laat ook zien dat het na de keuze voor het volgen van hierarchie in een
bepaalde dimensie niet mogelijk is om naar hierarchie in een andere dimensie
over te stappen. Deze manier van denken is gebruikt om Fig. 2 zo aan te
passen dat de nieuwe ordening wel rekening houdt met hiërarchische
dimensies. Het resultaat staat in Fig. 4. De linker kolom laat zien dat in het
heelal steeds deelgebieden ontstaan met toenemende differentiatie, te beginnen
met sterrenstelsels, gevolgd door zonnestelsels, zonnen, planeten en
ecosystemen. In de rechter kolom is voor enkele deeltjes aangegeven hoe de
complexiteit toeneemt door interne differentiatie. De middelste kolom toont dat
samenwerking of interne differentiatie de basis kan vormen voor een toename
van complexiteit van fysische deeltjes tot complexe organismen.
Samenvatting
249
Figuur 3. Drie fundamentele richtingen voor hiërarchie in de natuur. Pijlen naar boven
wijzen van organismen naar toenemende complexiteit in ecosystemen. Pijlen naar rechts
wijzen in de richting van organismen met toenemende complexiteit. Pijlen naar beneden
wijzen in de richting van hiërarchie in de interne organisatie van organismen.
Figuur 4 Een herschikking van de systemen in een ecologische hiërarchie waarbij is
uitgegaan van drie dimensies.
250
Samenvatting
Figure 5. De evolutie van de operatoren. De donkere lijn volgt het historische pad
waarlangs door middel van first-next possible closures de opeenvolgende operatoren zijn
gevormd. De systemen in de grijze kolommen zijn wel het gevolg van eerstvolgende
mogelijke closure, maar zijn strikt genomen geen operatoren omdat ze alleen een
cyclisch proces of een cyclische structuur (de ‘interface’) vertegenwoordigen en niet een
combinatie van beide.
Verklaring van afkortingen Fig. 5:
SAE (Structural Auto Evolution) - de eigenschap dat een operator zelfstandig het deel van zijn structuur
dat de informatie bevat, kan evolueren
SCI (Structural Copying of Information) - de eigenschap dat een operator zelfstandig de in zijn
hypercyclus vastgelegde informatie kan kopiëren door eenvoudigweg hiervan de structuur te kopiëren.
Deze informatie is vastgelegd in de elementen van de katalytische hypercyclus of in de neurale
hypercyclus.
Hypercycle - de first-next possible closure die is gebaseerd op cirkels bestaande uit cyclische processen
Interface - een first-next possible closure die een nieuwe ruimtelijke grens creëert.
CALM (Categorizing And Learning Module) - een minimale eenheid voor neuraal geheugen.
HMI (Hypercycle Mediating interface) - een first-next possible closure veroorzaakt een interface die
selectief interacties afschermt tussen de hypercyclus en de buitenwereld.
Multi-state - een operator gebaseerd op structurele en functionele first-next possible closure op basis
van operatoren van een closure-niveau lager.
Samenvatting
251
Een strikte rangorde
De hiërarchie van deeltjes in de middelste kolom van Figuur 4 vormt de basis
voor de theorie in dit proefschrift. In deze hiërarchie is in iedere volgende stap
een zelforganisatieproces te herkennen. Deze zelforganisatie leidt automatisch
tot een nieuw cyclisch proces en een nieuwe cyclische vorm. Door het sluiten
van deze cirkels ontstaat een nieuwe eenheid. Een kringvorm van dergelijke
interacties wordt in de systeemkunde algemeen aangeduid met het begrip
‘closure’, wat daarom ook in de hier ontwikkelde theorie is gebruikt. Closure
betekent in het Engels afronding of geslotenheid.
De theorie in dit proefschrift is gebaseerd op situaties in de natuur waar een
nieuw type deeltje en daaraan verbonden volgende closure automatisch wordt
gerealiseerd als de eerstvolgende mogelijkheid die de natuur toelaat. Deze
opeenvolging van twee deeltjes wordt hier aangeduid als ‘ first-next possible
closure’ (eerstvolgende mogelijke closure). Het gebruik van dit begrip maakt een
strikte ordening mogelijk en is daarom onmisbaar voor de hier uiteengezette
theorie. Het kijken naar de combinatie van gesloten processen en gesloten
vormen in combinatie met het principe van first-next possible closure leidt er toe
dat slechts één bijzondere groep systemen in het heelal aan dit principe voldoet.
Al deze systemen blijken via de eis van first-next possible closure met elkaar te
zijn verbonden in een grote afstammingsboom en vormen daarmee een strikt
hiërarchische deelverzameling van alle systemen in het heelal. Deze
deelverzameling omvat de fundamentele deeltjes, de hadronen, de atomen, de
moleculen, de prokaryote cellen, de eukaryote cellen, de pro- en eukaryote
meercelligen en de dieren.
Ieder element van deze verzameling is op basis van zijn closures van de
omgeving te onderscheiden als een individuele operationele eenheid. Dit is de
reden dat de verzameling van alle elementen in de hiërarchie in dit proefschrift
de naam ‘operatoren’ heeft gekregen.
De introductie van het concept operator biedt een belangrijke bijdrage aan het
samenbrengen van verschillende natuurwetenschapsgebieden. Het blijkt
namelijk dat in al deze wetenschapsgebieden een zelfde type bouwsteen, deze
zogenaamde operator, de basis vormt van de onderzochte systemen. Hiermee
krijgt het begrip operator een extra betekenis naast bestaande betekenissen in
het Engels, zoals een wiskundige bewerking of iemand die een apparaat
bedient, bijvoorbeeld een telefonist(e).
In Figuur 5 is met een donkere lijn de familiestamboom van de opeenvolgende
operatoren aangegeven. Deze hiërarchische ordening van alle operatoren is de
252
Samenvatting
‘operatorhiërarchie’ genoemd en is het resultaat van deze studie. De
operatorhiërarchie heeft, in tegenstelling tot alle eerder ontwikkelde
hiërarchische benaderingen van de natuur, de bijzondere eigenschap dat het
niet mogelijk is om één operator aan de hiërarchie toe te voegen of er uit te
breken zonder dat de hele structuur uiteen valt. Wat ook nieuw is aan de
operatorhiërarchie is dat de groep dieren een nieuwe naam heeft gekregen,
namelijk ‘memon’ (meervoud memons). Deze aanpassing maakt een helder
onderscheid mogelijk tussen organismen zonder en organismen met neuraal
netwerk. Zo is de mens bijvoorbeeld een organisme en een memon, terwijl een
plant alleen een organisme is. Bovendien is het woord memon in de toekomst
ook te gebruiken voor levende wezens die zijn gebaseerd op technische neurale
netwerken (de robot-achtige systemen of ook wel ‘technische memons’).
Nieuwe ordening verenigt disciplines
Het doel van dit proefschrift is om ordenende principes in de
natuurwetenschappen te vinden en deze toe te passen bij het bouwen van
bruggen tussen de disciplines. Hier volgt een beschrijving van de fundamentele
bijdragen die de operatortheorie kan leveren aan een brede integratie van de
exacte wetenschappen.
 De operatortheorie maakt het mogelijk om bestaande ecologische
hiërarchieën te verbeteren door rekening te houden met drie aspecten: de
hiërarchie van de operatoren, de interne organisatie van operatoren en
interacties tussen operatoren (zie Fig. 4). Het gemengd toepassen van deze
drie aspecten in hiërarchische benaderingen verstoort de logica daarvan.
 Gangbare definities van Darwiniaanse evolutie zijn sterk gefocused op
erfelijkheid, reproductieve diversificatie en selectie. De operatortheorie maakt
het mogelijk om de complexiteit van organismen toe te voegen aan de
evolutietheorie.
 De operatortheorie biedt een periodiek systeem voor de koppeling van alle
andere periodieke systemen (voor zover deze betrekking hebben op
operatoren).
 De operatortheorie biedt een nieuwe context voor het definiëren van leven.
Leven kan worden gedefinieerd als materie met de structuur van een
operator en met de minimale complexiteit van de (prokaryote) cellulaire
operator. Als consequentie hiervan zal ook de definitie van een organisme
als volgt moeten worden aangepast. Een organisme is een operator die
voldoet aan de definitie van leven. Sterfte treedt op als een irreversibel
vervalproces dat de organisatie van een organisme aantast, leidt tot het
verlies van de voor dit organisme kenmerkende meest complexe first-next
possible closure.
 Als mag worden aangenomen dat het principe van first-next possible closure
algemene geldigheid heeft in heel het heelal, dan moeten buitenaardse
Samenvatting





253
levensvormen op dezelfde first-next possible closures zijn gebaseerd als het
leven op aarde.
De operatorhiërarchie laat zien dat first-next possible closures kunnen
worden gevormd en ook weer uiteen kunnen vallen. De lange keten van
operatoren toont daarnaast aan dat het optreden van zelf-organisatie in de
vorm van first-next possible closure in thermodynamisch opzicht een
gunstige toestand moet zijn. Dit betekent dat de ontwikkelingen daarom altijd
voort zullen snellen in de richting van de volgende operator, tenzij locale
omstandigheden het zelforganizatieproces niet faciliteren.
De operatorhiërarchie biedt een nieuw raamwerk dat gerichte voorspellingen
mogelijk maakt van essentiële constructie-eigenschappen van toekomstige
operatoren. De operatortheorie voorspelt dat memons gebaseerd op
technische neurale netwerken de volgende stap vormen in de evolutie van de
operatoren. Onderbouwing door middel van de operatorhiërarchie geeft aan
een voorspelling een veel zwaarder gewicht dan aan extrapolaties gebaseerd
op bestaande trends in de ontwikkeling van robots. Bovendien kan op basis
van de operatorhiërarchie veel verder vooruit worden gekeken.
De operatorhiërarchie biedt een raamwerk om op een gestructureerde
manier de discussie te voeren of er meer is tussen hemel en aarde dan waar
de wetenschap zicht op heeft door aan te geven waar onbekende zaken zich
met grote waarschijnlijkheid niet bevinden.
Omdat de memon gewoon een willekeurige operator in de operatorhiërarchie
is, zijn er geen wetenschappelijke redenen om mensen te beschouwen als
het eindpunt van de evolutie of als een kroon op een vermoede schepping.
De operatortheorie laat zien, dat het wenselijk is om het begrip ‘meem’ voor
een idee in een informatiedrager zoals de hersenen (de “meme” senu
Dawkins) en daaraan verwante culturele replicatoren (melodieën, text, etc.)
aan te vullen met een extra interpretatie die wél een structurele parallel
vertoont met het begrip gen (Engels: gene). Dit kan door de meem ook te
laten verwijzen naar de structuur van het neurale netwerk van een memon.
Een meem krijgt dan de vorm van een serie codes die weergeeft welke
neuronen worden verbonden, welke verbindingen er bestaan tussen de
neuronen en welke sterkte de verbindingen hebben. In principe kan met een
dergelijke meem en de bijbehorende interface alle kennis van een memon
worden overgebracht op eventuele nakomelingen.
De volgende uitdaging
De operatorhiërarchie is een sterk vernieuwende theorie, maar desondanks niet
een volledig nieuwe benadering. De Franse paleontoloog, filosoof en Jezuïet
Pierre Teilhard de Chardin heeft halverwege de vorige eeuw een serieuze
poging ondernomen in dezelfde richting. Hij was op zoek naar een methode om
het christelijk geloof in overeenstemming te brengen met de evolutietheorie. In
254
Samenvatting
zijn benadering maakte hij onderscheid tussen enerzijds systemen die hij
‘gevormd’ en ‘gecentreerd’ noemde en anderzijds alle systemen zonder deze
eigenschappen. De aanduidingen ‘gevormd’ en ‘gecentreerd’ vertonen veel
gelijkenis met de structurele en functionele closures die bepalend zijn voor
operatoren. Teilhard de Chardin was er sterk van overtuigd dat alleen het
ordenen van gevormde en gecentreerde systemen inzicht in de evolutie kon
opleveren. Hoe bewonderenswaardig ook, de theorie van Teilhad de Chardin
was te weinig onderbouwd om wetenschappelijk geaccepteerd te worden.
Bovendien werd zijn benadering door onderzoekers gewantrouwd omdat zijn
hiërarchie eindigde met de eenwording van de mensheid met god in het
zogenaamde punt omega. De operator theorie gebruikt kernen uit het
gedachtegoed van Teilhard de Chardin en breidt deze uit met moderne inzichten
en een gedetailleerde systeemkundige onderbouwing. Bovendien maakt de
operatortheorie aannemelijk dat de evolutie van operatoren juist geen vast
eindpunt heeft.
Omdat de operatortheorie nieuw is, zijn er nog veel uitdagingen om deze theorie
verder te verfijnen. Eerst en voor alles is het noodzakelijk om de regels die de
basis vormen voor de kolommen in Figuur 5 beter te leren begrijpen. Een beter
begrip van deze regels zal het namelijk mogelijk maken om voorspellingen over
de volgende stadia in de evolutie aan te scherpen en op basis hiervan
nauwkeuriger uitspraken te doen over toekomstige operatoren. Dit zal
wetenschappers en filosofen nieuwe mogelijkheden bieden om op verkenning te
gaan in een domein waarvan men tot voor kort dacht dat het ontoegankelijk was
voor een natuurwetenschappelijke benadering: de toekomst van de evolutie.
History and acknowledgements
“We rarely forget that which has made a deep impression on our minds.”
(Tryon Edwards)
256
History and acknowledgment
The evolution of an idea
The operator hierarchy started developing while I was working as a postdoc at
the Vrije Universiteit from September 1992 to August 1993 after my PhD in
ecotoxicology. The project was initiated by the “Netherlands Integrated Soil
Research Programme” and involved writing an integration study of 21 PhD
projects on soil pollution. The project started with writing fact sheets for all 21
projects. Based on this and other information, a proposal for theoretic integration
entitled “Ecotoxicology across the levels of ecological organisation” was sent to
the advisory committee in mid-1993. This draft already hinted at using
dimensions as an organising principle, stating that “all stress on ecological
systems is a combination of the effects on the creation and destruction of
‘matter’ and ‘information’ in systems”. Around this time Bas Kooijman allowed
me to read a first draft of his book “Dynamic energy budgets in biological
systems”. Near the end of 1993, the project’s advisory committee received a
second draft report. Later that year I read Kauffman’s book “The Origins of
Order”, and the basic idea for the operator hierarchy was born during the winter
of 1993-1994. Developing the operator hierarchy required studying many new
things ranging from quarks to neural networks, which took a while. Around this
time Bart Happel introduced me to the astonishing world of technical, modular,
recurrent neural networks and the fractal inner space they show. Figure 1 shows
one of the first drawings of the operator hierarchy and illustrates just how much
work had yet to be done.
Figure 1. This figure illustrates one of the first drawings of the operator hierarchy (spring
1994). Note that every level shows the same number of elements.
History and acknowledgments
257
In August 1994, a final report on the integration project was sent to the advisory
committee with the title “Application of a hierarchical concept of ecosystem
organisation in ecotoxicology”. The report included a part about the dynamic
balance of toxicants in organisms and a part about targets, buffering and
plasticity and the processes of assessment, diagnosis and prognosis. The
advisory committee responded with disappointment because they had hoped for
indications of “What is an acceptable toxin level?”. Instead, the report dealt with:
When, Why, How, Related to what, and even So What? The report introduced
the basic ideas of the ‘operator’ and ‘closure’: “Using closure of operations as a
borderline between two hierarchical levels in the information evolution, the
organisation shows four levels: the sub-atomary level, the level of atoms and
molecules, the organic level, and the level of the consciousness”. Importantly, it
was concluded that “… the repeating of an evolutionary pattern within the
hierarchical levels appeared to be so eminent, that we have given the resulting
hierarchy a central position in this study”. It was a great help to discuss the
operator hierarchy with Cajo ter Braak in January 1995. His comments included
a list of 92 suggestions and questions in the manuscript’s margins. By this time,
the operator hierarchy had evolved into the scheme in Figure 2.
Figure 2. The operator hierarchy as it looked in January 1995. The numbers in circles
refer to Cajo ter Braak’s remarks. Note that the scheme now assumes a variable number
of steps per level, but that complexity increase is still linear within levels.
258
History and acknowledgment
In 1996, while working at NERI (Denmark,) I submitted the Lego-bricks paper.
After two years it was sent back unreviewed. Klaus Skovbo Jensen then
introduced me to Eigen’s hypercycles. In 1997, I was inspired by Lars Witting
who kindly gave me a draft of his book “A General Theory of Evolution”. In 1998,
Hans Løkke gave me the green light for three months of writing time to restyle
the ecotoxicology manuscript. But after these months of hard work -- during
which I stayed in a NERI guesthouse -- it appeared that I had again
underestimated the complexity and quantity of the work. In 1999, the paper
about the Lego bricks was published in World Futures, the journal of general
evolution. And later that year, a new job came along at Alterra as the head of the
functional biodiversity team.
During the last months of 2001, Diedel Kornet introduced me to her idea of “slots
in state space” and gave me an interesting scheme of the natural levels
including elementary particles, atoms, molecules, free-living cells, multicellular
organisms and interbreeding communities. This scheme also contained the
principle of internal organisation. In 2001, the Acta Biotheoretica paper was
published. The prediction of future operators then awoke the media’s warm
interest because of the question: “Will robots dominate humanity?” This resulted
in a range of interviews for local and national newspapers and for radio and
television. Suddenly, living in two worlds proved too complex. I decided to create
more time for the operator hierarchy. This proved a fruitful strategy. The
Analysing Hierarchy paper was published in Biological Reviews in 2008. In
2009, the ‘definition of life paper’ was accepted by Foundations of Science (to be
published in 2010). In June 2009, Rolf Hoekstra invited me to a meeting in his
office with Eörs Szathmáry. After a short presentation of the operator hierarchy’s
major ideas, Eörs made two very valuable comments. As far as he knew, the
idea of first-next possible closure was a unique approach. He also indicated that
using only interactions as a basis for first-next possible closure in the ‘three
dimensions picture’ was presumably too restrictive. This was a wise remark. I
immediately adapted the figure. Based on the accepted publications, I
considered the possibility of a PhD. When speaking to Nico van Straalen about
this plan, he advised me to contact Hub Zwart of the Radboud University
Nijmegen. This resulted in a very pleasant contact and two month of writing time
to finalise the project.
History and acknowledgments
259
Acknowledgements
“Without friends no one would choose to live,
though he had all other goods.”
(Aristotle)
During the many years that the ‘project’ of the operator hierarchy was underway,
it received contributions from a broad range of people who were either directly
involved in the project or whom I talked with at universities, at institutions, at
meetings and at conferences. I am grateful to all those who have inspired me. A
special word of thanks should go to the following people for their more than
average contribution. Nico van Straalen from the Vrije Universiteit in Amsterdam
has supported me theoretically and practically during all the years in which the
NISRP study metamorphosed into the operator hierarchy. Cees van Gestel did a
wonderful job guiding me almost daily during the NISRP ecotoxicology study. I
would also like to thank the 21+ researchers involved in the NISRP study:
Maarten Nederlof, Marca Schrap, Sandra Plette, Peter Doelman, Martien
Janssen, Marianne Donkers, Henk Siepel, Matty Berg, Joop Hermens, Hans de
Kruijf, Joke van Wensum, Erik van Capelleveen, Jan Kammenga, Carla Roghair,
Cees van de Guchte, Jos Nootenboom, Lieuwe Haanstra, Jan-Willem
Kamerman, Trudie Crommentuijn, Hans Vonk, Patrick van Beelen, Wim Ma,
Carl Denneman, Hélène Loonen, Angelique Belfroid and others. Their
discoveries formed the fundaments of this thesis. I am grateful to the advisory
committee of the NIRSP for their feedback: Nico van Straalen, Cees van Gestel,
Herman Eijsackers, Hans de Kruijf, Willem van Riemsdijk, Anton Stortelder and
Willem Salomons
It was a great honour to receive and a pleasure to study an early draft of Bas
Kooijman’s book on dynamic energy budgets. I still regard this work as the Ayers
Rock of ecotoxicology. I am grateful for general discussions about the operator
hierarchy and for input from a practical systems perspective from Doeke
Doeksen and Gert Faaken. Cajo ter Braak commented in great detail on an early
version of the operator hierarchy and offered much appreciated personal
support. Bart Happel from Leiden University introduced me to the field of
technical neural networks.
At the National Environmental Research Institute in Silkeborg (NERI, Denmark),
numerous people have supported the development of the operator hierarchy and
have become dear friends, notably Hans Løkke, who arranged three months of
writing time, Jorgen and Lisa Axelsen, Christian and Mette Kjaer, Christian
Damgaard and Inger Møller.
260
History and acknowledgment
At Alterra I am most grateful to Henk Siepel for many years of practical and
theoretical support of the project. Jeff Baldwin critically read the concept text of
the hypercycle website and Wiert Dijkkamp made it accessible on the internet.
Rob Hengeveld helped me organise and submit my book manuscript “The
Modular Evolution of Matter”, which, unfortunately, was rejected. Hans Peter
Koelewijn offered new insights into genetics and suggested various excellent
books with up-to-date theories.
For commenting on various drafts of papers, I thank Henk Siepel, Herbert Prins,
Hans Peter Koelewijn, Leen Moraal, Ton Stumpel and Dick Melman. Rolf
Hoekstra has been involved in the operator hierarchy project in various ways,
most notably in supporting the Analyzing Hierarchy paper and by inviting me for
a meeting with Eörs Szathmáry. Clément Vidal and John Smart are thanked for
organising the Evo-Devo conference in Paris. I thank Peter Winiwarter, who died
all too early, for his friendship, his inspiring website and his plan to include the
operator hierarchy in his book. Furthermore, I would like to thank all members of
the Blauwe Kamer Koor for their much appreciated and warm friendship, and
especially Albert Ballast, Marnix van der Meer, Peter Schippers and Marjolein
Kikkert, all of whom contributed to the operator hierarchy through discussions
and valuable insights.
I thank Claire Hengeveld for the challenging job of improving the English of the
Biological Reviews paper. Liewien Koster contributed by commenting on the
definition of life paper. I thank Palmyre Oomen and Ronny Desmet for
information about Whitehead. During the preparation of the thesis, Hub Zwart of
the Radboud University Nijmegen contributed by arranging two months of writing
time, by offering inspiring new points of view and excellent suggestions on the
style of the final texts. Peter Griffith’s corrections of the English have greatly
improved the texts of the unpublished chapters of this thesis. Reena BakkerDhaliwal helped improve the English of the ‘stellingen’. Last but not least, I am
grateful for the assistance of Desiree Hoonhout at the Vrije Universiteit
Amsterdam, Birgit Sørensen at the NERI and Nelja Schönthal, Suze Landman,
Liesbeth Steenhuisen and Leona van Eck at Alterra. Florentine Jagers op
Akkerhuis greatly improved the quality of the draft text of the summary, the
readability of which was subsequently tested by Ivo Laros and Leen Moraal. The
craftsmanship of Henny Michel transformed my plain texts to a nice ‘Alterra
Scientific Contribution’, the cover of which was given shape by Martin Jansen.
Of course, this thesis would never have been produced without all the moments
of joy shared with family and friends. I would need many more pages to thank
everyone who enriched our lives at home, in the Floriade, during Christmas
dinners and the Ekocup, at the Koninginnemarkt, during different swampeturene,
on Livø, in Gammle Ry and on Westerbakken, in the Gorges du Tarn, in Death
Valley, Las Vegas, San Francisco, at meetings and conferences in Warsaw,
History and acknowledgments
261
Odense, Paris, Winchester and Budapest, during the coffee breaks, during
rehearsals and weekends with the BKK, while having our yearly Roghorst
barbeque, during the City of Insects festival, in Xanten, while listening to the
‘wiew wiew’ of stone owls, when counting deer in de Kop van Schouwen, during
the World of Music Festival, etc.
Above all, I am grateful to Florentine, Marrit, Stijn and Inge. You have been and
still are the best friends I could ever wish for. Thank you for your practical and
mental support, for tolerating my periods of absentmindedness and for never
giving up teasing me about ‘het evolutie gedoe’ (‘the evolution stuff’).
262
History and acknowledgment
Curriculum vitae
“We are the accidental result of an unplanned process …
the fragile result of an enormous concatenation of improbabilities,
not the predictable product of any definite process.”
(Stephen Jay Gould)
264
Curriculum vitae
My work involves using system approaches and evolutionary theory to analyse
practical problems in biology, ecology and ecotoxicology. My work has
contributed to studies concerning the effects of agricultural practices on
biodiversity, the modeling of insecticide side-effects, the biodiversity impacts of
climate change, invasive species, forest biodiversity, the biodiversity of the soil
ecosystem, advisory work in relation to the habitats directive and birds directive,
biodiversity monitoring, trait-based ecology, etc. As a general basis for the
understanding and analysis of systems, I developed the operator, which is
explained in this thesis and on my website hypercycle.nl.
Period
Degree
1978 - 1986
MSc (Cum Laude)
Locations and topics
Wageningen Agricultural University: Entomology,
Nematology, Animal physiology
1988 - 1993
PhD
Wageningen Agricultural University: Ecotoxicology
(Prof. J.H. Koeman)
1992 - 1993
Post Doc
Vrije Universiteit Amsterdam: Ecotoxicology
(Prof. N.M. van Straalen,)
1995 - 1995
Researcher
Centrum voor Landbouw en Milieu
1996 - 1999
Guest scientist
NERI, Silkeborg, Denmark (Dr H. Løkke)
1999 - 2010
Senior scientist
Alterra, Wageningen UR, Nederland (Prof. H. Siepel)
2009
Senior scientist
Radboud University, Nijmegen, (Prof. H. Zwart)
Publications
First author (related to the operator hierarchy)
Jagers op Akkerhuis G.A.J.M. (2010, in press). Explaining the origin of life is not enough for a
definition of life. Foundations of Science.
Jagers op Akkerhuis G.A.J.M. (2010, in press). Towards a hierarchic definition of life, the organism,
and death. Foundations of Science.
Jagers op Akkerhuis G.A.J.M. (2009). The Operator Hierarchy: Evolution Beyond Genes. Abstract of
a poster presentation at ECAL 2009, Budapest.
Jagers op Akkerhuis G.A.J.M. (2008). Analysing hierarchy in the organisation of biological and
physical systems. Biological reviews 83: 1-12
Jagers op Akkerhuis, G.A.J.M. (2001). Extrapolating a hierarchy of building block systems towards
future neural network organisms. Acta Biotheoretica 49: 171-189.
Jagers op Akkerhuis, G.A.J.M. (2001), Essaybundel Kennisdisseminatie: Technische intelligentie is
volgend stadium in de evolutie (Dutch)
Jagers op Akkerhuis, G.A.J.M, Damgaard C. (1999): Using ‘resource dominance’ to explain and
predict evolutionary success. Oikos 87:609-614.
Jagers op Akkerhuis, G.A.J.M., van Straalen N.M. (1998): Operators, the Lego-bricks of nature,
evolutionary transitions from fermions to neural networks. World Futures, The journal of
general evolution (53: 329-345).
Curriculum vitae
265
First author (other)
Jagers op Akkerhuis G.A.J.M., Dimmers W.J., van Vliet P.C.J., Goedhart P.W., Martakis G.F.P., de
Goede R.G.M. (2008). Evaluating the use of gel-based sub-sampling for assessing
responses of terrestrial microarthropods (Collembola and Acari) to different slurry
applications and organic matter contents. Applied soil ecology 38: 239 – 248
Jagers op Akkerhuis G.A.J.M., Dimmers W.J., van Kats R.J.M., van der Hoef M., Gaasenbeek
C.P.H., Schouls L.M., van de Pol I., Borgsteede F.H.M. (2003): Observations on Lyme
disease, tick abundance and the diversity of small mammals in different habitats in the
Netherlands and implications for the 'inflation' and 'dilution' effect approaches. Proc. Exper.
Appl. Entomol. 14: 31-38.
Jagers op Akkerhuis, G.A.J.M.; Siepel, H. (2002). Effects of anthelmintics on manure-eating
arthropods. Vakbl. Natuurbeheer 41 (2002), May spec. Issue: 40
Jagers op Akkerhuis, G.A.J.M., Seidelin, N., Kjær, C. (1999): Are we analysing Knockdown in the
right way? How independence of the knockdown-recovery process from mortality may affect
measures for behavioural effects in pesticide bioassays. Pestic. Sci. (55: 62-68)
Jagers op Akkerhuis, G.A.J.M., Damgaard, C., Kjær, C., Elmegaard, N. (1999): Comparison of the
toxicity of dimethoate and cypermethrin in the laboratory and the field when applying the
same bioassay. Environ. Toxicol. Chem. 18: 2379-2385
Jagers op Akkerhuis, G.A.J.M., Kjær, C., Damgaard, C., N., E. (1999): Temperature Dependent,
Time-dose-effect model for pesticide effects on growing, herbivorous arthropods; bioassays
with dimethoate and cypermethrin. Environ. Toxicol. Chem. 18: 2370-2378
Jagers op Akkerhuis, G.A.J.M., Axelsen, J.A., Kjær, C. (1998): Towards predicting pesticide
deposition from plant phenology; a study in spring barley. Pestic. Sci. 53: 252-262.
Jagers op Akkerhuis, G.A.J.M., Rossing, W.A.H., Piet, G., Everts, J.W. (1997): Water depletion, an
important cause of mortality in females of the spider Oedotohorax apicatus after treatment
with deltamethrin: a simulation study. Pestic. Biochem. Physiol. 58: 63-76.
Jagers op Akkerhuis, G.A.J.M., Westerhof, R., Van Straalen, N.M., Koeman, J.H. (1995): Water
balance, respiration and immobilisation in relation to Deltamethrin poisoning and physical
conditions in the epigeal spider Oedothorax apicatus. Pestic. Sci. 44: 123-130.
Jagers op Akkerhuis, G.A.J.M. (1993): Physical conditions affecting pyrethroid toxicity in arthropods.
PhD thesis, Wageningen Agricultural University,197 pp.
Jagers op Akkerhuis, G.A.J.M. (1993): Walking behaviour and population density of adult linyphiid
spiders in relation to minimizing the plot size in short term pesticide studies with pyrethroid
insecticides., pp. 163-171, Environ. Poll. 80.
Jagers op Akkerhuis, G.A.J.M., Van der Voet, H. (1992): A dose effect relationship for the effect of
deltamethrin on a linyphiid spider population in winter wheat., pp. 114-121, Arch. Environ.
Cont. Toxicol. 22.
Jagers op Akkerhuis, G.A.J.M., Hamers, T.H.M. (1992): Substrate dependent bioavailability of
deltamethrin for the epigeal spider Oedothorax apicatus (Blackwall)(Aranaea, Erigonidae).,
pp. 59-68, Pestic. Sci.36.
Jagers op Akkerhuis, G.A.J.M., de Ley, F., Zwetsloot, H.J.C., Ponge, J.-F., Brussaard, L. (1988): Soil
microarthropods (Acari and Collembola) in two crop rotations on a heavy marine clay soil.,
pp. 175-202, Rev. Ecol. Biol. Sol. 25.
Jagers op Akkerhuis, G.A.J.M., Sabelis, M., Tjallingii, W.F. (1985): Ultrastructure of chemoreceptors
on the pedipalps and first tarsi of Phytoseiulus persimilis. Experim. Applied Acarology 1: 235251.
266
Curriculum vitae
co-author
Mulder C., H.A. Den Hollander, J.A. Vonk, A.G. Rossberg, G.A.J.M. Jagers op Akkerhuis and G.W.
Yeates (2009). Soil resource supply influences faunal size–specific distributions in natural
food webs. Naturwissenschaften 96:813-26.
Rutgers, M.; Schouten, A.J.; Bloem, J.; Eekeren, N.J.M. van; Goede, R.G.M. de; Jagers op
Akkerhuis, G.A.J.M.; Wal, A. van der; Mulder, C.; Brussaard, L.; Breure, A.M. (2009).
Biological measurements in a nationwide soil monitoring network. European Journal of Soil
Science 60: 820 - 832.
van der Wal, A., Geerts, R.H.E.M., Korevaar, H., Schouten, A.J., Jagers op Akkerhuis, G.A.J.M.,
Rutgers, M., Mulder, C. (2009). Dissimilar response of plant and soil biota communities to
long-term nutrient addition in grasslands. Biology and Fertility of Soils 45 (6), pp. 663-667
Rutgers, M., Mulder, C., Schouten, A.J., Bloem, J., Bogte, J.J., Breure, A.M., Brussaard, L., De
Goede, R.G.M., Faber, J.H., Jagers op Akkerhuis, G.A.J.M., Keidel, H., Korthals, G.,
Smeding, F.W., Ter Berg, C., Van Eekeren, N. (2008) Soil ecosystem profiling in the
Netherlands with ten references for biological soil quality. Report 607604009, RIVM,
Bilthoven, The Netherlands.
Wielinga, P.R., Gaasenbeek, C., Fonville, M., de Boer, A., de Vries, A., Dimmers, W., Jagers op
Akkerhuis, G., Schouls, L., Borgsteede, F., van der Giessen, J.W.B. (2006). Longitudinal
analysis of tick densities and Borrelia, Anaplasma, and Ehrlichia infections of Ixodes ricinus
ticks in different habitat areas in the Netherlands. Applied-and-Environmental-Microbiology
72: 7594-7601
Good, J.A. and Jagers op Akkerhuis G.A.J.M. 2005. Interpreting breeding habitat from the presence
of adults when Staphilinidae (Coleoptera) are used in biodiversity assessments. Bull. Ir.
Biogeog. Soc. 29: 194-212.
Bloem, J. T. Schouten, W. Didden, G. Jagers op Akkerhuis, H. Keidel, M. Rutgers, T. Breure. 2004.
Measuring soil biodiversity: experiences, impediments and research needs. Proceedings of
the OECD expert meeting on soil erosion and soil biodiversity indicators, 25-28 March 2003,
Rome, Italy. OECD, Paris, p. 109-129.
Moraal, L.G., G.A.J.M. Jagers op Akkerhuis, D.C. van der Werf, 2003. In: A.J.H. van Vliet, M.
Grutters, J.A. den Dulk & R.S. de Groot (eds): Challenging times; towards an operational
system for monitoring, modelling, and forecasting of phenological changes and their socioeconomic impacts; abstracts. Wageningen,
Wageningen-UR-Environmental
Systems
Analyses Groep, pp 85.
Elmegaard, N., Jagers op Akkerhuis, G.A.J.M. (2000): Safety factors in pesticide risk assessment.
(NERI Technical Report, No. 325, 60 pp.).
Van Wensem, J., Jagers op Akkerhuis, G.A.J.M., Van Straalen, N.M. (1991): Effects of the fungicide
Triphenyltin hydroxide on soil fauna mediated litter decomposition. Pestic. Sci. 32: 307-316.
Aukema, B., Berg, J.H.J.v.d., Leopold, A., Jagers op Akkerhuis, G.A.J.M. (1990): A method for
testing the toxicity of residues of pesticides on a standardized substrate to erigonid and
linyphiid spiders., pp. 76-80, J. Appl. Ent. 109.
Arpin, P., Jagers op Akkerhuis, G.A.J.M., Ponge, J.-F. (1988): Morphometric variability in Clarkus
papillatus (Bastian, 1865) Jairajpuri, 1970 in relation to humus type and season. Revue
Nématol. 11: 149-158.
Curriculum vitae
267
author (Dutch)
Jagers op Akkerhuis, G.A.J.M., W. Dimmers, M. Maslak, N. van Eekeren en A.J. Schouten (2009).
Microarthropoden als indicatoren van de kwaliteit van landbouwgronden. De invloed van
mest en biologische bedrijfsvoering op de bodem. Alterra-rapport 1985.
Jagers op Akkerhuis, G.A.J.M., A.H.P. Stumpel, W.J. Dimmers, B. Verboom (2008). Inventarisatie
van
de
flora
en
fauna
in
2008
in
‘de
Landschapszone’,
onderdeel
van
het
ontwikkelingsgebied ‘de Waalsprong, gemeente Nijmegen. Alterra-rapport 1749, 64 pp.
Jagers op Akkerhuis, G.A.J.M. (2008). Inventarisatie van compensatiegebied voor de Steenuil rond
‘de Waalsprong’, gemeente Nijmegen. Alterra-rapport 1782, 33 pp.
Jagers op Akkerhuis, G.A.J.M., A.H.P. Stumpel, B. Verboom, H. van Blitterswijk (2008). Aanvullende
inventarisaties van vogels met vaste nesten, amfibieën en vleermuizen in de wijken Groot
Oosterhout, Laauwik en Citadel in het ontwikkelingsgebied de Waalsprong te Nijmegen.
Alterra-rapport 1732, 37 pp.
Jagers op Akkerhuis, G.A.J.M., Moraal L.G., Veerkamp M.T., Bijlsma R.J., Vorst O. & van Dort K.
(2007). De rol van doodhoutspots voor de biodiversiteit van het bos – Veldonderzoek naar
de rol van doodhoutspots de vestiging van zeldzame insecten, paddenstoelen en mossen.
Alterra-rapport 1435. 62 pp.
Jagers op Akkerhuis G.A.J.M., S.M.J. Wijdeven, L.G. Moraal, M.T. Veerkamp & R.J. Bijlsma (2006).
Dood hout en biodiversiteit. Een literatuurstudie naar het voorkomen van dood hout in de
Nederlandse bossen en het belang ervan voor de duurzame instandhouding van
geleedpotigen, paddenstoelen en mossen. Alterra rapport 1320.
Jagers op Akkerhuis G.A.J.M., L.G. Moraal, M.T. Veerkamp, R.J. Bijlsma & S.M.J. Wijdeven (2006).
Dood hout en biodiversiteit. Vakblad Natuur Bos Landschap 3: 20-23.
Jagers op Akkerhuis, G.A.J.M. Groot Bruinderink, G.W.T.A., Lammertsma, D.R., Kuipers, H. (2006).
Biodiversiteit en de ecologische hoofdstructuur: een studie naar de verdeling van soorten
over Nederland en de dekking van hun leefgebieden door de ecologische hoofdstructuur.
Alterra rapport 1319.
Jagers op Akkerhuis, G.A.J.M., Lammertsma, D.R., & Groot Bruinderink, G.W.T.A. (2004). Een
meetnet voor de Nederlandse soortdiversiteit; nationale en internationale afspraken,
bestaande meetnetten en een aanbeveling voor een meetnet soortdiversiteit. Alterra-rapport
(Int. rep. 1063). Alterra.
Jagers op Akkerhuis, G., L. Moraal, J. Burgers, R. van Kats, D. Lammertsma, W. Dimmers, A.
Noordam., B. Aukema, F. Bianchi, W. van Wingerden (2004). Biodiversiteit in het agrarisch
landschap. Ekoland 2004-6, 20-22.
Jagers op Akkerhuis, G.A.J.M., Lammertsma, D.R., Martakis, G.F.P., & Berg, A. van den (2004).
Functioneel ecologische voorwaarden voor hotspots van biodiversiteit; verspreiding van
cryptobiota met zweefvliegen (Syrphidae) als voorbeeld. Alterra-rapport (Int. rep. 937).
Alterra.
Jagers op Akkerhuis, G.A.J.M., W.J. Dimmers, R.J.M. van Kats, M. van der Hoef, C.P.H.
Gaasenbeek, L.M. Schouls, I. van der Pol & F.H.M. Borgsteede (2003). Observations on
Lyme disease, tick abundance and the diversity of small mammals in different habitats in the
Netherlands and implications for the 'inflation' and 'dilution' effect approaches. Proc. Exp.
Appl. Entomol. 14: 31-38.
Jagers op Akkerhuis G., L. Moraal, J. Burgers, R. van Kats, D. Lammertsma, W. Dimmers, A.
Noordam, B. Aukema, F. Bianchi, W. van Wingerden. Biodiversiteit in het agrarisch
landschap. Manipulatie van populaties nuttige insecten. Ekoland 6-2004: 20-22.
268
Curriculum vitae
Jagers op Akkerhuis, G.A.J.M., W.J. Dimmers, R.J.M. van Kats, M. van der Hoef, C.P.H.
Gaasenbeek, L.M. Schouls, I. van der Pol & F.H.M. Borgsteede, 2003. Observations on
Lyme disease, tick abundance and the diversity of small mammals in different habitats in the
Netherlands and implications for the 'inflation' and 'dilution' effect approaches. Proc. Exp.
Appl. Entomol. 14: 31-38.
Jagers op Akkerhuis, G.A.J.M., W.J. Dimmers, R.J.M. van Kats, M. van der Hoef, C.P.H.
Gaasenbeek, F.H.M. Borgsteede, L.M. Schouls, I. van der Pol, 2003. Meer muizen, meer
teken, meer ziekte van Lyme. Infectieziekten Bulletin 14: 170-171.
co-author (Dutch)
Groot Bruinderink, G.W.T.A., L. van Breukelen. Damherten en reeën in het natuurreservaat de Kop
van Schouwen. Inventarisaties. Alterra-rapport 1933, 2009, 84 pp.
Groot Bruinderink, G.W.T.A., D.R. Lammertsma, G.A.J.M. Jagers op Akkerhuis, W. Ozinga, A.H.P.
Stumpel, J.M. Baveco en R.W. de Waal. Ex ante evaluatie van maatwerk beheer van wilde
zwijnen. Alterra-rapport 1944.
Bakker, G., G.A.J.M. Jagers op Akkerhuis, I.E. Hoving (2009). Hoogwatergeul Veessen-Wapenveld.
Inschatting van de gevolgen van tijdelijke inundaties op bodemstructuur, bodemleven en
grasland. Alterra-rapport 1890, 47 pp.
Rutgers, M., Jagers op Akkerhuis, G.A.J.M., Bloem, J., Schouten, A.J., Breure, A.M. (2009)
Prioritaire gebieden in de Kaderrichtlijn Bodem: belang van bodembiodiversiteit en
ecosysteemdiensten. Rapport 607370001, RIVM, Bilthoven.
Faber, J.H., G.A.J.M. Jagers op Akkerhuis, J. Bloem, J. Lahr, W.H. Diemont & L.C. Braat.
Ecosysteemdiensten en transities in bodemgebruik; Maatregelen ter verbetering van
biologische bodemkwaliteit. Alterra-rapport 1813, 2009, 150 pp.
Groot Bruinderink, G.W.T.A., D.R. Lammertsma, A.T. Kuiters, R.G.M. Kwak, R.J.H.G. Henkens &
G.A.J.M. Jagers op Akkerhuis, m.m.v. RIZA: J. Th. Vulink & P. Cornelissen. Verbinding
Oostvaardersplassen – Hollandse Hout; Onderdeel van de uitvoering van het ICMO-advies.
Alterra-rapport 1595, 2007, 87 pp.
Groot Bruinderink, G.W.T.A., R.J. Bijlsma, M.A.K. Bleeker, H. Esselink, G.A.J.M. Jagers op
Akkerhuis, D.R. Lammertsma, F.G.W.A. Ottburg, A.H.P. Stumpel, W.C.E.P. Verberk & E.J.
Weeda Pilot Leefgebiedplan Laagveenmoeras. Een ecologische uitwerking van het concept
leefgebiedbenadering. Alterra-rapport 1548, 2007, 85 pp.
Weeda, E.J., W.A. Ozinga & G.A.J.M. Jagers op Akkerhuis Diversiteit hoog houden; bouwstenen
voor een geïntegreerd natuurbeheer. Alterra-rapport 1418, 2007, 246 pp.
Booij, K, J. Lahr, G.A.J.M. Jagers op Akkerhuis: Hoe belangrijk is het agrarische gebied voor de
insectendiversiteit?. Entomologische Berichten 2007, 67: 184-186.
Lahr, J., K.Booij, D. Lammertsma, G. Jagers op Akkerhuis: Nederlandse biodiversiteit. Hoe
belangrijk is het agrarisch gebied? Landschap2007, 3: 109-115.
Kemmers, R.H., J. Bloem, J.H. Faber & G.A.J.M. Jagers op Akkerhuis Bodemkwaliteit en
bodembiodiversiteit bij natuurontwikkeling op voormalige landbouwgronden. Alterra-rapport
1523, 2007, 50 pp.
Vosman, B., H. Baveco, E. den Belder, J. Bloem, K. Booij, G. Jagers op Akkerhuis, J. Lahr, J.
Postma, K. Verloop en J. Faber. Agrobiodiversiteit; Kansen voor een duurzame landbouw.
PRI-rapport 165, 2007, 72 pp.
Borgsteede, F., C. Gaasenbeek, A. de Boer, J. Dijkstra, G. Jagers op Akkerhuis, W. Dimmers, P.
Wielinga, M. Fonville, L. Schouls, I. van der Pol, J. van der Giessen. ASG-rapport
Curriculum vitae
269
06/1100015/mak, 13 pp. Het verloop van tekenpopulaties en de besmetting van teken met
Borrelia en Ehrlichia. Resultaten van onderzoek in de periode 2000-2005.
Wolf, R.J.A.M., W.J. Dimmers, P.W.F.M. Hommel, G.A.J. Jagers op Akkerhuis, J.G. Vrielink & R.W.
de Waal Bekalking en toevoegen van nutriënten. Evaluatie van de effecten op het
bosecosysteem. Een veldonderzoek naar vegetatie, humus en bodemfauna. Alterra-rapport
1337.5, 2006, 42 pp, € 0
Lahr, J., G.A.J.M. Jagers op Akkerhuis, C.J.H. Booij, D.R. Lammertsma & J.J.C. van der Pol
Bepaling van het belang van het agrarisch gebied voor de biodiversiteit in Nederland. Een
haalbaarheidsstudie. Alterra rapport 2005, 72 pp
Moraal, L.G., G.A.J.M. Jagers op Akkerhuis, J. Burgers, W.J. Dimmers, D.R. Lammertsma, R.J.M.
van Kats, G.F.P. Martakis, Th. Heijerman & J. Poutsma Oriënterend onderzoek naar
Geleedpotigen in liggend dood hout van Zomereik en Grove den. Alerra-rapport 1101, 2005,
69 pp
Haveman, R., J. Burgers, W.J. Dimmers, H.P.J. Huiskes, G.A.J.M. Jagers op Akkerhuis, R.J.M. van
Kats, D.R. Lammertsma & G.F.P. Martakis Evertebraten in faunaranden en natuurbraak; een
detailstudie in Noordoost-Groningen, gepubliceerd: 01 Feb 2005, 64 pp
Moraal, L.G., G.A.J.M. Jagers op Akkerhuis, H. Siepel, M.J. Schelhaas & G.F.P. Martakis
Verschuivingen van insectenplagen bij bomen sinds 1946 in relatie met klimaatverandering;
Met aandacht voor de effecten van stikstofdepositie, vochtstress, bossamenstelling en
bosbeheer. Alterra-rapport 856,: 2004, 68 pp
Bloem, J. T. Schouten, W. Didden, G. Jagers op Akkerhuis, H. Keidel, M. Rutgers, T. Breure. 2004.
Measuring soil biodiversity: experiences, impediments and research needs. Proceedings of
the OECD expert meeting on soil erosion and soil biodiversity indicators, 25-28 March 2003,
Rome, Italy. OECD, Paris, p. 109-129.
Goede de R.G.M., P.C.J van Vliet, B. van der Stelt, F.P.M. Verhoeven, E.J.M. Temminghoff, J.
Bloem, W.J. Dimmers, G.A.J.M., Jagers op Akkerhuis, L. Brussaard, W.H. van Riemsdijk,
2003. Verantwoorde toepassing van rundermest in graslandbodems. SKB rapport nr. SV-411
pp.48.
Moraal, L.G., A.F.M. van Hees, G.F.P. Martakis, I.T.M. Jorritsma, G.A.J.M. Jagers op Akkerhuis,
2003. Een karakterisering van bosbiotopen op basis van eigenschappen van geleedpotigen;
resultaten van een enquete. Alterra-report 783, 72 blz.
A.M. Breure, M. Rutgers, J. Bloem, L. Brussaard, W.A.M. Didden, G.A.J.M. Jagers op Akkerhuis, C.
Mulder, A.J. Schouten, H.J. van Wijnen, 2003. Ecologische kwaliteit van de bodem. RIVM
Bilthoven, RIVM rapport 607604005, pp. 32.
Wingerden, W.K.R.E. van, M. Nijssen, P.A. Slim, J. Burgers, R.J.M. van Kats, H. F. van Dobben,
A.P. Noordam, G.F.P. Martakis, H. Esselink & G.A.J.M. Jagers op Akkerhuis Grazers in
Vlielands duin. Evaluatie van runderbegrazing in duinvalleien op Vlieland; deel 2: onderzoek
in 2001. Alterra-rapport 626: 2003, 100 pp.
Wingerden, W.K.R.E. van, M. Nijssen, P.A. Slim, J. Burgers, G.J.A.M. Jagers op Akkerhuis, A.P.
Noordam, G.F.P. Martakis, H. Esselink, W.J. Dimmers & R.J.M. van Kats Evaluatie van
zeven jaar runderbegrazing in duinvalleien op Vlieland. Alterra-rapport 375, 2001, 102 pp
Faber, J.H., G.A.J.M. Jagers op Akkerhuis, J. Burgers, B. Aukema, R.J.M. van Kats, G.F.P.
Martakis, D.R. Lammertsma & A.P. Noordam Ongewervelde fauna van ontkleide
uiterwaarden; Eindrapport. Alterra-rapport 372, 2003, 122 pp.
270
Curriculum vitae
Glossary
“Accurate definitions improve the precision and communication of science.
Sloppy definitions lead to the development of sloppy theory.
A lack of definitions leads to no science at all.”
(Gerard Jagers op Akkerhuis).
272
Glossary
The operator theory represents a new point of view. This implies that a number
of new concepts had to be introduced and defined. Furthermore, as a
consequence of the logic of the operator theory certain existing concepts were
re-evaluated and sometimes had to be re-defined. Additionally, there exist
various concepts in science that play a more general role in the context of the
operator hierarchy and for which it is useful to have a short description at hand.
For all these reasons it was considered appropriate to offer the readers of this
thesis an overview of concepts that play an important role in the context of the
operator theory.
Analogy: Analogy implies a mapping of knowledge from one domain (the base)
to another (the target) such that a system of relations that holds among the base
objects also holds among the target objects (Gentner & Jeziorski 1993)
Building block: I sometimes use the word 'building block' or 'unit systems' for
indicating the particles that in the operator hierarchy are named operators. I
always have in mind a system in which the elements create a physical and
functional unity on the basis of first-next possible closures.
Catalytic hypercycle: see Hypercycle
Closure: In the context of the operator hierarchy, closure relates to a state of
matter that is the result of a self-organisation process and that shows a closed
topology with respect to structure, process or the combination of both. Closure
as a state shows a close association with underlying mechanisms because it
also is used to refer to the closing process producing the closed state.
Closure level: In the context of the operator hierarchy, every first-next possible
closure adds one closure level.
Closure, structural and functional: Functional closure is defined as a closed
cycle of processes that does not cause a physical mediating layer. Specific
forms of functional closure are the catalytic reactions involved in an autocatalytic
set, the pion exchange between protons and neutrons in the atom nucleus and
the plasma strands connecting the cells of multicellulars. In contrast, the
definition of structural closure demands that interactions create a physically
closed topology. In the operator hierarchy, the focus is on a special form of
structural closure, namely the structural closure that creates a physical boundary
mediating a contained, first-next possible, hypercyclic process. Specific
examples of structural closure in the operator hierarchy are the electron shell
around the atom nucleus, the cell membrane of a single cell, and the connected
cell membranes of multicellular organisms and the sensory interface of the
memon.
Glossary
273
Closure dimension: A closure dimension focuses on similarity in structural
closure, functional closure or, when present, a combination of both. The naming
of a closure dimension is directed after the most complex organisational property
of the system as allowed by structural or functional closure or, when possible,
the combination of both. The following closure dimensions are acknowledged in
the operator hierarchy: 1. the interface dimension, introduced by the
fundamental particles, 2. the hypercyclic dimension, introduced by the quarkgluon plasma, 3. the multi-state dimension, introduced by the hadrons, 4. the
hypercycle mediating interface dimension (HMI), introduced by the atoms, 5. the
structural copying of information dimension (SCI), introduced by the cell, 6. the
structural auto-evolution dimension, introduced by the hardwired memon
Complexity (functional): I regard as the functional complexity of an entity 'all
forms of internal or external interactive processes that are supported by the
internal organisation of the entity’. Both a large, badly organised structure and a
minute, well-integrated structure may allow an entity to show the same functional
complexity.
Complexity: Complexity is a concept that refers to our inability to give a
straightforward description of the properties of an entity. Complexity relates to
the number and inter-relatedness of the elements of an entity and to the patterns
and sub-patterns in its states and dynamics.
Complexity of a closure dimension: The complexity of a closure dimension is
based on the number of major transitions that were required to construct it.
According to the operator theory, the complexities of the closure dimensions that
are known to exist are: 1 for the fundamental particles, 2 for the first-next
possible closure creating hypercyclic interaction systems, 3 for the hadrons, 4
for the atoms, 5 for the cell and 6 for the memon.
Complexity of operators: The operator hierarchy measures the complexity of
an operator by means of the number of first-next possible closure levels that
underlie its construction.
Confinement: Confinement binds quarks in hadrons in a condensation-like
process. It is the result of the strong force, also called the color force, which is
conveyed by the exchange of gluons between quarks. At higher
temperatures/energies the quarks become increasingly heavy and the
confinement relatively weaker, which is referred to as the asymptotic freedom of
the quarks.
Corrupt hierarchy: If, in any hierarchy, one or more steps/layers do not comply
with the hierarchy rule used for creating the ranking, I consider this a corrupt
hierarchy. For example the hierarchy 1, 2, 4, 5, 6 ... is corrupt in relation to the
274
Glossary
rule that for every N, the next element in the ranking must be exactly N+1 (it
lacks the 3). For the same reason the ranking 1, 2, 3, 3.2457, 4 ... is a corrupt
hierarchy (the 3.2457 is superfluous).
Dissipative system: The organisation of a dissipative system results from
selforganisation processes caused by the degradation of free energy gradients
in the environment. Examples of dissipative systems are waves (degrading wind
energy), whirlwinds (degrading a pressure gradient in the atmosphere) and
organisms (degrading a chemical and/or radiation energy gradient). As a
thermodynamic equilibrium implies a random distribution of elements and
interactions, the organisation of elements in a dissipative system is regarded as
to be 'far from equilibrium'. Dissipative systems do not contradict with the laws of
thermodynamics, because any increase in organisation inside a dissipative
system (which implies a decrease in entropy) is driven by a free energy gradient
and causes at least as much entropy in the system as is lost during the
organisation process.
Efficiency: I regard efficiency as the lowest resource use, when two processes
are compared leading to a functionally equivalent product or achieving a
functionally equivalent goal.
Elementary closure: An elementary closure represents the lowest complexity
realisation of its type with respect to the system’s structure, process, or when
possible, the combination of both. Elaboration and repetition of the same closure
type does not affect the elementary closure.
Elementary particle: In physics, the concepts of elementary particles and
fundamental forces are used to indicate matter particles and force carrying
particles of the standard model, respectively. The operator hierarchy suggests
that one may also regard all operators as elementary particles with respect to
their highest level first-next possible closure. Defined this way, only the particles
of the standard model are truly elementary (which I prefer to consider
fundamental) in the sense that they presumably show no substructure of smaller
particles.
Emergent property: An emergent property is a group-property that results from
interactions between entities that individually do not show this property. For
example water flow and waves are emergent properties of interacting water
molecules (a single molecule cannot form a wave). Used in this general way,
emergent properties arise from almost any interaction between separate
elements.
Entropy: Entropy is the opposite of free energy. Because in a closed system the
average free energy can never increase, the entropy can only increase. An
Glossary
275
increase in entropy corresponds with the natural pathway of systems from a high
to a low energy state. Examples are the reaction of chemicals to low energy
products, the transformation of light to heat, the falling of particles in a
gravitational or electric field and the change of a system towards the state that
shows the highest number of possible microstates. For closed systems
Boltzmann has shown that the chance on a decrease in entropy becomes
infinitesimally small for systems consisting of many elements. It is an important
aspect of nature that in local parts of open systems entropy may decrease, the
local system becoming more organised. This does not violate the laws of
thermodynamics, as long as the entropy decrease in the local system is driven
by a free energy gradient.
Evolution: Any process based on the following aspects: 1. the production of
entities the structure of which differs from that of their precessors (productive
diversification), and 2. structure-dependent performance of the entities causing a
gradient in their evolutionary success (the cause of selection). Evolutionary
dynamics can be observed at all levels of the operator hierarchy (e.g.
elementary particles, prions and viruses, organisms, technical memons) but also
things that are not operators may show evolution, e.g. strings of computer code,
bee colonies, or neuron states produced in relation to the invention and weighing
of different scenarios in a decision making process. The evolution of neuron
states precedes that of tools (cars, windmills and the like) with which it coevolves.
Evolutionary success: Evolutionary success is a measure that depends on the
relative performance of entities with respect to: 1. the stability of their internal
organisation, 2. the stability of their functioning in interactions with other entities
and/or forces, 3. their capacity for productive diversification, which in relation to
operators ultimately includes the capacity to produce the following first-next
possible closure. Without the latter, evolution cannot proceed towards higher
level operators.
The evolutionary success of dissipative systems, such as living beings, is related
to the relative success with respect to resource dominance, when compared to
other systems that take part in the process. The longer an organism exists and
the better it functions (which may include reproduction), the more resources it
(and its offspring) will use/dominate. Organisms which at a given moment are
the best resource dominators, may rapidly lose from others which get slowly
better over time, whilst both strategies will be beaten by organisms that can
increase the rate of their improvement, especially if they can increase the
acceleration of this rate. Strength in competition as well as strength in
cooperation may increase evolutionary success. The duration of observations
determines which of the above aspects of evolutionary success can be
observed.
276
Glossary
A higher evolutionary success not necessarily implies a more complex structure.
For example, in dynamic environments with a high random mortality, simple,
rapidly reproducing entities will prevail.
Exhaustive closure: To assist in identifying the new type of elementary closure,
I suggest identifying the ‘exhaustive closure’ in the system at the one lower
closure level. A system shows exhaustive closure, if there is no remaining
potential for elaborating the structural and dynamic aspects of the elementary
closure type, because any further development will cause the construction of a
new elementary closure type. While the elementary closure of the system
remains the same, the state of exhaustive closure allows for a new functionality.
First-next possible closure: First-next possible closure is defined as follows:
given any system A that shows first-next possible closure, the next first-next
possible closure creates the least complex system type above A that shows a
new type of elementary closure based on A and, when required, any highest
level system type possible below A that shows first-next possible closure. This
definition is inherently recursive because a system showing first-next possible
closure produces a more complex system showing first-next possible closure.
Because a first-next possible closure is always built from systems showing a
preceding first-next possible closure, the recursive definition does not lead to
logical loops.
Force: A force is the result of an interaction. For example the repulsion between
electrons (electromagnetism) is caused by the interaction of two electrons by
means of the exchange of virtual photons. Likewise, the exchange of gluons
causes the attractive force between quarks. In principle the equating of force
and interaction works al all levels of the operator hierarchy. For example the
interaction between a predator and its prey can be regarded as a ‘predation
force’.
Fundamental particle: A physical particle that shows no substructure of smaller
particles. In the operator hierarchy: the system showing the lowest level firstnext possible closure (see also elementary particle).
Hardwiring: This concept is used to describe a neural architecture that is based
on physical 'wires' between neurons, for example in the shape of dendrites. The
difference with 'softwiring' is that softwired connections between neurons do not
exist as physical wires, but as entries in computer memory telling which neuron
is connected to which other neurons and what is the strength of the connection.
Hierarchy: Hierarchy assumes a ranking of entities. Formally, elements in a
hierarchy meet the following two demands (Simon 1973): Irreflexivity: A can
never hold a hierarchical position below itself. Transitivity or non-transitivity: if A
Glossary
277
has a lower hierarchical position than B, and B has a lower position than C, then
A does have, or does not have a lower position than C. From irreflexivity and
(non-)transitivity follows Antisymmetry: if A has a lower hierarchical position than
B, then B cannot hold a lower position than A. An example of a transitive
hierarchical relationship is: A is ancestor of B, B is ancestor of C. In this case A
is also ancestor of C. An example of a non-transitive hierarchy is: A is a child of
B, B is a child of C. In this case A is not a child of C.
Bottom-up hierarchies arise when the behaviour of lower level elements
integrates them in larger assemblies (e.g. when atoms create a molecule). Topdown hierarchies may have different origins, for example 1. when social
agreements define a top figure (marshal/manager) and allow it to direct groups
of (now) lower level figures (soldiers/workers) or 2. when after the formation of a
functional unit, internal differentiation leads to internal hierarchies of functional
relationships (see also 'Penultimate level').
Hierarchical layer: Operators and pre-operator interaction systems that are
based on the same primary system and do not yet represent the next major
transition, are regarded as to belong to the same hierarchical layer.
Hypercycle: The use of this concept in the literature generally refers to the
enzymatic hypercycle as discussed by Eigen and Schuster (1977). With respect
to the operator hierarchy, the concept is used for any second order cyclic
process that is related to a first-next possible closure.
Information and complexity: Information has a structural and a semantic
aspect. The structural aspect relates to the number of elements and the different
states they can code for (e.g. Shannon 1948). The semantic aspect relates to
the interpretation of a coding. In principle, there is no limit to the semantic
complexity of a message, because even a simple coding may relate to a highly
complex interpretation. The functional elements of a coding (the ‘bites’,
'characters', 'genes', 'symbols', etc.) normally show an intermediate complexity.
Too low complexity would allow too little differences between the elements. Too
high complexity would make the coding/decoding difficult.
Intelligence: In the context of the operator theory intelligence is an emergent
property brought about by the functioning of an operator with a complex enough
memic architecture. Intelligence is associated with the autonomous capacities to
observe, learn, make internal representations, predict, evaluate and act.
Intelligence requires sensors for contacting the world and a neural network for
creating internal multiple-channel representations, the latter offering a natural
source for association and creativity. Intelligence shows a gradual increase with
increasing neural complexity. For intentional behaviour, the representations
must be available for the evaluation of which actions are most effective for
278
Glossary
reaching the mental concept of a certain goal. Intelligence, therefore, has much
to do with the art of choosing appropriate aspects for evaluation in multi-factorial
problems.
Interaction: An interaction is an event that involves at least two entities and
causes a change in one or more parameters of both systems. This excludes as
interactions all events in which one element remains unchanged. For example, if
someone listens to a radio program, this seems an interaction between the
person speaking on the radio and you. But the person speaking is not affected in
this process, and therefore it is not an interaction according to the above
definition.
Interaction system: An interaction system consists of interacting operators but
lacks the closure(s) that would allow it to be regarded as an operator. The world
is swarming with interaction systems, including footballs, stars, cars, water,
populations, society, companies, etc. Of course also operators consist of
interacting elements. Yet, like Orwell made his pigs advocate that 'all animals
are equal, but some are more equal than others', I advocate that all systems are
interaction systems, but that the operators show such special interactions, that it
is wise to regard them as 'more equal than others'. Accordingly I assign them to
a special subset. Although they play an important role in the operator hierarchy,
the pre-operator hypercyclic sets and interfaces are not operators, but
interaction systems.
Life; as a state of matter: Life is a concept that relates to all matter that shows
the construction of operators, and that has the minimum complexity of the
cellular operator. Subsets of living entities can be defined as autocatalytic life
(uni- and multicellular plants, fungi, sponges, protozoa, etc.) or memic life
(including all memic operators, regardless whether they are based on a cellular
or a technical construction).
Life; meaning of: The meaning of life is in the first place a personal subject.
Besides this, the operator theory offers two general ways of thinking about the
meaning of life. The first meaning is associated with the functioning of organisms
as individual operators. Every organism has to stay alive and memic organisms
additionally care about their daily satisfaction. Survival is a quite objective
criterion. The validation of daily satisfaction depends on the individual judgment
of what is 'satisfying'. The second meaning of life is associated with the
contribution of all operators to evolution. On the one hand, this implies an
involuntary act. Whatever you do, you are always part of the large process that
is driven by entropy production. On the other hand, increasing insight in the
overall structure of evolution may have the implication that a person decides to
realise a directed contribution to evolution, for example by constructing a
technical memon
Glossary
279
Living: Living can be defined as the dynamic activity of systems that comply
with the definition of life. A bacterium, for example, may or may not be living,
depending on whether it is active or frozen/dried.
Major transition: Any first-next possible closure creating a primary system is
regarded as a major transition (not to be confused with the major evolutionary
transitions of Maynard Smith and Szathmáry, 1995a, 1995b)
Meme: Dawkins (The selfish gene, 1976) defines the meme as: '... new
replicators, which I called memes to distinguish them from genes, can propagate
themselves from brain to brain, from brain to book, from book to brain, from
brain to computer, from computer to computer.' Dawkins' memes are not
structural analogues of genes, because any pattern of replicable information in
brains, books, etc. could be a meme, whilst not any pattern of replicable
information in cells is regarded a gene. To solve this, I propose the use of
different meme concepts: 1. Abstract meme (A-meme): abstractions
communicated between memons (melodies, stories, and theories), 2. Physical
memes (P-meme): physical models of thoughts (written language, artwork,
buildings), 3. Functional meme (F-meme): the actual neural network, or part of it,
harboring information, 4. Coding meme (C-meme): C-memes code for neural
network architecture just as genes code for catalytic molecules.
Minor transition: Any first-next possible closure creating a system that shows
an already existing closure dimension, is regarded a minor transition.
Model: I regard as a model any entity that represents another entity. This may
involve mental models that represent parts of our environment, scaled physical
models that represent larger or smaller physical or mental originals, and
simulation models that represent aspects of real world systems.
Modularity: Modularity refers to the construction of larger systems from small
functional and/or structural subassemblies or 'modules'. Herbert Simon
illustrated the evolutionary merits of modularity in the parable of the
watchmakers Tempus and Hora, making watches consisting of 1000 parts each.
Tempus made his watches bit by bit. When disturbed, he had to redo the entire
assembling process. Hora put his watches together from subassemblies of 10
parts each. 10 subassemblies were put together in a larger subassembly and 10
larger subassemblies constituted the whole watch. Hence, when Hora was
disturbed he only lost the specific subassembly-job he was involved in. Simon
calculated that if customers disturbed Hora and Tempus on average once in
hundred assembly operations, Tempus would need about four thousand times
longer to assemble a watch.
280
Glossary
Natural systems: I regard as natural systems all elements of the universe of
which the physical existence allows us to interact with it and describe static
and/or dynamic properties (qualities and quantities) in a consistent way.
More/better measurements should lead to an increasing detail of our
representation of the system. Kragh (Aktuel Naturvidenskab 1999) has said this
in the following way. 'It is difficult to understand, how measurements of important
scientific parameters normally can become more and more precise, and how
knowledge shows a general consolidation and refining, if these measurements
and knowledge do not concern something that really exists in nature'. In the
context of the operator theory, natural systems include the neural configuration
and neural dynamics that 'carry' a thought, for example of the number pi, but do
not include the word 'thoughts' or the number ‘pi’.
Operator: An operator is defined as a system type that shows first-next possible
closure with a closure dimension of at least 3. This implies that operators can
also be defined as those systems showing first-next possible closure that are not
pre-operator interaction systems. The set of the operators unifies all particles,
irrespective whether these are physical particles or organisms. All operators that
are currently known to exist are: the hadron, the atom, the molecule, the
(bacterial) cell, the eukaryote cell, the bacterial and eukaryote multicellular and
the neural network organism, called the ‘memon’ in the operator hierarchy.
Operator hierarchy: The operator hierarchy is the structuring of all operators
and pre-operator interaction systems on the basis of first-next possible closure
and closure dimensions.
Operator hypothesis: The operator hypothesis states that the fundamental
limitations imposed by first-next possible closure act as a strict mould for particle
evolution.
Operator theory: The operator theory is the theoretical framework that offers
the context for the operator hierarchy and the operator hypothesis.
Penultimate level (Law of the branching growth of the …): This law by
Turchin (1977) states that: '...after the formation, through variation and selection,
of a control system C, controlling a number of subsystems Si, the Si will tend to
multiply and differentiate. The reason is that only after the formation of a
mechanism controlling the Si it becomes useful to increase the variety of the Si.
Pre-operator interaction systems: Systems, caused by first-next possible
closure, that selectively create an interface or a 2nd order cyclic interaction (the
‘hypercycle’) as closure dimension, are called pre-operator interaction systems.
Glossary
281
Primary system: A system showing a new closure dimension is regarded a
primary system.
Quanta of evolution: Quantification normally involves simple variables such as
length, weight, color (wavelength), etc. A more complex variable to quantify is
the evolutionary complexity of a system, for which quanta of evolution have been
proposed (Turchin 1995, Heylighen, Joslyn and Turchin 1995). The operator
hierarchy uses first-next possible closure steps for creating a strict system of
quanta of organisation.
Recursion of definition: A definition is recursive if it explains some entities Xn
in terms of X. The operator hierarchy is recursively defined, because lower level
operators are used to define higher level operators. Note that a recursive
definition does not suffer from a logical loop.
Resource: A resource is something a living operator needs to enable certain
aspects of its existence.
Resource dominance: Resource dominance is defined as the capacity of
organisms to gain dominance over resources. There are different aspects of
resource dominance that are of evolutionary importance. First, there is the
amount of resources that an organism can dominate by degrading them. After
degradation they are no longer available for other organisms. Second, there are
resources that an organism may dominate, without that it uses them. Again they
are no longer available for other organisms. There exist various resource
dominance strategies by which an operator can increase its evolutionary
success.
Softwiring: In softwired memons, the neural network is based on computer files
that keep track of the neurons involved, the interactions they show with other
neurons, and the types and connection strengths of the interactions.
System: The system concept relates to entities which can be endowed groupwise functionality because of past, present or future interactions and/or
relationships. The entities that by their relationships define the group are
considered as the elements of the system. The system concept applies to the
universe or a subset of it (the latter including mental states and via these, all
aspects of our imagination). Frequently, the entities and their interactions are
chosen in such a way that they define, include or relate to some kind of physical
system limit.
System type of closed systems: Two systems that are created by the same
first-next possible closure are considered to be of the same system type.
282
Glossary
Tool: In the context of the operator hierarchy I define as a tool 'any idea or
physical object that a memic operator may use to help it manipulating the world
to its liking'. Consequently there are mental tools, such as mathematics, and
physical tools, such as a stick, a hammer, a car, a house, a factory and a
computer. It is interesting to realise that the operator hierarchy predicts that the
next level memons are not based on cells but on technical hardware.
Consequently, the process of reproduction (giving birth), that is so important for
the evolution of organisms based on one or more cells, will be replaced by the
process of producing an 'offspring-tool'.
Topology: Topology studies qualitative aspects of geometrical structures. In
stead of asking how big a certain thing is, it focuses on other aspects, such as:
does it have any holes in it; is it connected together, or can it be separated into
parts.
Transition: If a system changes from state A to state B this can be regarded as
a transition of the system between these states. For any particular transition the
property that is going to be looked at as well as the states that can be
recognised have to be defined.
Type: A type is used as a grouping of entities according to a common property.
Using types of types leads to the creation of a type-hierarchy in which lowerlevel types show properties that are grouped into higher-level types. For
example the higher-plant-type and the animal-type can be grouped within the
multicellular-organism-type. Frequently used types in relation to the operator
hierarchy are: 1. process type: a comparable aspect that can be recognised in
the dynamics of different entities. 2. System type (general): the type of a system
is defined by comparable aspects with respect to a. the rules involved in the
selection of elements, b. the properties defining the elements and c. the
relationships that may exist between the elements, and 3. Closure type: a
closure type is defined by the presence of comparable aspects with respect to
the type of elements and the type of mechanisms causing the closure.
Index
Index
283
284
Index
actual occasions 21
anti-color 38
anti-symmetry 17
arrow of complexity hypothesis 205, 206
astrobiology 204
atom 26, 41
autocatalysis 43
autopoiesis 22, 29, 123
baryon 39
Big Bang 79
Big History 177, 232
biological systems 103
biomatrix 23
body factory 94
boson 70
bottom-up construction 19
building block systems 66, 79
butterfly-effect 11
catalytic hypercycle 43
categorising and learning module (CALM)
52, 71, 83
cell 27
cell; first cell 29, 45
cell; omnis cellula e cellula 28
centeredness 21
centripetal science 10
chaos theory 11
chemosystem 113
chemoton theory 23
Chinese boxes 19, 65
chloroplast 28
circular relationships; importance of 104
closure 11, 23, 33, 164
closure dimensions; (in-)dependence of 86,
96, 167, 169
closure type 85
closure; and hypercycle dimension 167
closure; both process and state 34, 36
closure; elementary closure 164
closure; exhaustive closure 166
closure; first-next possible closure 33, 35,
106, 163, 165
closure; hypercycle mediating interface
dimension 168
closure; interface dimension 167
closure; multi-state dimension 168
closure; structural and functional 34
closure; structural autocopying of information
dimension 168
closure; structural auto-evolution dimension
168
closure; transitive, cyclic, surjective 24
color and anti-color 38
color charge and color force 27, 38
community; conceptual part of ecosystem 14
compartmentation 66
complexity; inward complexification 32
complexity; levels in nature 12, 17, 235
complexity; of operators i.r.t. resource
dominance 151
complexity; of operators 157
complexity; yardstick for 34
compound object 109
construction sequence vs. evolutionary
sequence 117
constructivism 20
cosmic onion 65
covalent bonds 42
death; definition of 136
definition as language filter 127
definition of death 7
definition of life 7, 127, 204
definition of organism 7, 128
density dependence 159
development acceleration 208
developmental stage of the universe 113,
116
DICE 19, 112
dimensions for analysing hierarchy 15, 179
discontinuity between system states 20
ecological hierarchy (renewed) 183, 238
ecological hierarchy (standard) 11, 13, 182,
236
ecosystem vs. chemosystem 113
ecotoxicology 12, 13
efficiency 157
electron microscope 29
embryo; from separate cells to the organism
49
emergence 65, 84, 144
endosymbiont hypothesis 29
endosymbiosis; complex forms of 30
endosymbiosis; step towards eukaryotic cell
46
energy gradients; degradation of 105, 209
evolution of evolvability 209
evolution theory; algorithm connecting
hypercycles 76
evolution theory; autocatalysis as driving
force 186
evolution theory; future of evolution 88-95
evolution theory; general laws for operators
75
evolution theory; interdisciplinary 64, 80, 81,
108, 181, 190, 234
evolution theory; levels of complexity 235
evolution theory; religious viewpoint of future
97
evolution theory; role of mankind 96
evolution theory; struggle for life, survival of
the fittest 151
evolutionary sequence vs. construction
sequence 117
evolutionary success 151
fermion 70
fitness 151
formedness 21
fundamental particles 25
future operators 7, 94
gap junctions 50
general system theory 17
generatio spontanae 28
gluon 70
gluon; quark gluon exchange 37
grand unifying theory 25
great chain of being 175
Index
hadron 39
hardwired memon 53, 88, 127, 186
hierarchic order; theory for 18, 104
hierarchical layer 170
hierarchy 17
hierarchy; bottom-up construction 19
hierarchy; constitutive vs aggregative 104,
105
hierarchy; in the organisation of nature 7, 59,
101, 182, 183
hierarchy; inconsistent, mixed or corrupt 14,
18, 66, 101
hierarchy; internal/somatic 105
hierarchy; linear 101
hierarchy; problems and controversy 18, 119
hierarchy; rules 17
hierarchy; standard or renewed ecological
hierarchy 11, 13, 182, 183, 236,
238
hierarchy; strict (no missing links) 24, 34, 59
hierarchy; three dimensions for analysing
hierarchy 15, 179, 202, 203, 237
hierarchy; transitive, irreflexive, asymmetrical
104
holism 19
holism; insect colonies, organs and brains 19
holon 22
hypercycle formation 66
hypercycle mediating interface dimension 83
hypercycle 23, 43
integrating theory 7
intelligence; based on hypercyclic neural
networks 96
intelligence; computer intelligence 97
interaction cycle; first order 37
interaction cycle; second order 37
interaction group 109, 110
interaction system 66, 67, 79, 106, 171
interaction system; complexity levels 113
interface; activation 54
interface; perception 54
interface; sensory 53
internal organisation 67, 102, 110, 115, 184
internal organisation; stability 58
irreflexivity 17
karyos; development 30
knowledge explosion and knowledge
explosion crisis 7, 10, 175, 232
knowledge; definition 53
law of the branching growth of the
penultimate level 22, 184
law of time and chaos 58, 209
lepton 70
lichens 51
life and death 136
life as we don't know it or extraterrestrial 134,
204
life; container, metabolism and genetic
program 123
life; criteria for organism 23, 128
life; definition based on operator hierarchy
125, 147
life; definition of life 123, 127, 204
285
life; discussion of operator-based definition
143, 144, 146
life; hierarchical levels of life 126, 129-132,
147, 148
life; meanings in different contexts 127
life; relationship with metabolism 134
life; relationship with reproduction 123
life; theory of life 124
life; use of facultative aspects in definitions
123
maintenance; of the cell 46
major evolutionary transitions 24, 82
major transition (in the operator theory) 106,
168
mediating effect of higher level system 117
meme 7, 54, 205
meme; of abstract, physical or coding type
55, 85
memon 54
memon; hardwired 53, 81, 88, 127, 135, 186
memon; HMI 92
memon; SCI 88
memon; technical 97
meson 39
metasystem transition 22
minicolumns in the brain 72
minor transition 83, 106, 168
mitochondria; anoxybiontic origin 30
mitochondrion 28
modularity 22, 208
molecule; as multi-atom 69
Muller’s ratchet 189
multi-atom 42
multicellular organism 48
multicellularity; and environmental conditions
50
multi-memon 86
mycorrhiza 51
neural network; efficiency 205
neural network; hardwired 53, 81, 88, 127,
135, 186
neural network; prewired 53
neural network; recurrent 52
neutron 26
neutron; half-life 40
nexus 21
non-equilibrium thermodynamics 187
nucleus of atom 40
Ockham's razor 175
operator (in operator theory) 12, 67, 80, 169
operator hierarchy 124, 169, 170
operator hierarchy; higher order logic 36, 56,
73, 76, 81, 108, 181, 240
operator hierarchy; predictions of future
operators 85, 86, 204
operator hierarchy; sensu stricto vs. sensu
lato 107
operator hierarchy; use of closure
dimensions 200
operator hierarchy; use of elementary
closure 200
operator hierarchy; use of first-next possible
closure 200
286
Index
operator hypothesis 171
operator theory; falsifiability 55, 57, 206-213,
210, 211
operator type; systems that change operator
type 112
operator vs. interaction system 106
operator; first of the row 82
operator; multistage 82
organisation; importance of circular patterns
104
organisation; of set of self maintaining
molecules 43
organism; criteria for living organism 23
organism; definition of the organism 128
Ouroboros 176, 232
particle physics 25
particle theory and particle concepts 25, 203
penultimate level 22
periodic table of operators 73
periodic table of periodic tables 7, 203
periodic table of the elements 10, 26, 184
periodic table; hadrons 184
periodic table; quarks 184
physical systems 103
pion exchange 26, 40
plasmodesmata 50
point Omega 33
pre-operator interaction system 169
primary system 168
property lists; inutility for definition of life 137
proton 26
quantum of evolution/organisation 22, 65,
124
quark 27, 70
quark; bare 27
rationale for extrapolating evolution trends 95
reductionism 19
regress; of system complexity 21
replayed tape metaphor 11
reproduction; of the cell 46
resource dominance strategies 151, 152,
154, 155, 156
resource; definition 152
Russian dolls (babushkas) 19
Scala Naturae 10, 175
self-organisation 11, 105
self-organisation; cyclical 7
sensory interface 53
stochiometry 23
structural autocopying of information
dimension 83
superstrings 27
system of interest; identification 114
system organisation; four step approach to
103
system; biological/physical 103
system; definition 162
system; elements of 103
system; interaction system 66
system; internal organisation 114
system; stability 105
targets for stressors 16
theory of everything 25
theory of life 124
thermodynamics and non-equilibrium theory
187
thermodynamics and system stability 57, 58
time 18
timeline 177
transition state 110, 125
transitivity 17, 18, 104
tree of life 10, 107, 109, 185
tree of life; levels of structural organisation in
30, 203
true natural units 32
type; definition 163
unifying concepts in science 192-196
universal timeline 177, 178
universe; developmental stage 113, 116
vehicle; for memon 54, 95
worlds within worlds approach 19, 65